content
stringlengths 1
15.9M
|
---|
\section{Introduction}
Primordial black holes (PBHs) arise naturally in most cosmologies.
Perhaps the least speculative mechanism for PBH formation comes from
the collapse of overdense regions in the primordial density
fluctuations that gave rise to structure in the universe
\cite{carrhawking}. Thus PBHs carry information of an epoch about
which we know comparatively little, and are a very useful tool for
restricting theories of the very early universe, especially within the
context of an inflationary scenario. In
this paradigm the density spectrum is a consequence of the
quantum fluctuations of the inflaton field, and can in principle be
calculated given an underlying model. Thus, the non-observation of
the by-products or of the effects of the energy density of these
PBHs constrains the underlying microscopic theory.
The simplest bound that can be extracted from PBH formation is generated
by insisting that $\Omega_{PBH}\leq 1$. Other bounds may be derived by
studying the consequences of their evaporation. Given that black
holes evaporate at a rate proportional to their inverse mass
squared~\cite{hawking},
the phenomenological relevance of a PBH will depend upon its initial
mass. Smaller mass PBHs ($10^9 < M_{bh} < 10^{13}$~g) will alter the
heavy elements abundances \cite{nuc} as well a distort the microwave
background \cite{micro}, whereas PBHs with larger masses will affect
the diffuse gamma-ray background \cite{diffuse,mac2,macreport}.
The net evaporation spectrum from a collection of PBHs will depend
on the initial mass distribution, which in turn depends upon the
probability distribution for the density fluctuations.
Prior to the emergence of the COBE data, bounds from PBHs were
calculated assuming a Harrison-Zel'dovich spectrum $|\delta_k|^2 \propto k$,
with an unknown normalization. This lead to the
famous ``Page-Hawking'' bound and all
its subsequent improvements \cite{diffuse,mac2}.
However, assuming that the distribution is Gaussian,
we can now relate the mass variance at the
time of formation with the mass variance at large scales today, if
we know the (model-dependent) power spectrum.
None the less, bounds on $n$ have been derived within the class of
models where the power spectrum is given by a power
law $|\delta_k|^2 \propto k^n$ over the scales of
interest \cite{carrlidsey,cgl,green,kim1,kim2}.
Indeed, in this way it follows that for the
scale-invariant Harrison-Zel'dovich spectrum PBHs have too small a number
density to be of any astrophysical significance. However, observations (see
references in \cite{cgl}) now seem to favor a tilted blue spectrum
with $n>1$ (in CDM models) with more power at smaller scales,
and thus the bounds derived from black hole evaporation can be used to
constrain the tilt of the
spectrum.
In this paper we revisit the aforementioned bounds in light of
recent calculations which indicate that a spectrum of primordial black
hole masses are produced through near critical gravitational collapse.
As was pointed
out by Jedamzik and Niemeyer \cite{JZI}, if PBH formation is a result
of a critical phenomena, then the initial mass function will be quite
different then what was expected from the classic calculation of Carr
\cite{carr}. In particular, the PBH mass formed at a given epoch is no
longer necessarily proportional to the horizon mass. The resulting
difference in the initial mass function leads to new bounds, which is
the main thrust of this paper. In particular, we revisit the density bounds
and the bounds derived from the diffuse gamma-ray observations.
We will first derive bounds on $n$ in the class of models where
the power spectrum is a power law over the scales of interest.
We then relax this assumption and instead place
bounds on the mass variance at the formation epoch.\footnote{As will
be discussed later, PBH formation is dominated by the earliest
formation epoch if they form via critical collapse.} These bounds can
then be applied to a chosen model by extrapolating the
variance today to the formation epoch according to the appropriate
power spectrum.
Before continuing to the body of this work, we would like to point out
that primordial black holes have also played a large role in
attempting to explain various data. PBHs can serve as significant
cosmological flux sources for all particle species via Hawking
radiation\cite{hawking}. Thus it is very tempting to postulate that
present-day observed particle fluxes of unknown origin are a consequence
of PBH evaporation. However, to predict these fluxes,
or model them realistically, we need to know the mass distribution of black
holes, since their emission spectra are determined by their
temperature (or inverse mass). If we assume that the black holes of
interest were formed from initial density inhomogeneities generated in
an inflationary scenario (which is usually assumed), then the black holes
are either tremendously over-abundant or completely negligible. To get a
phenomenologically interesting quantity of PBHs thus requires an extreme
fine-tuning, as will be demonstrated below. Succinctly, this fine-tuning
arises because the PBH number density is an extremely rapidly varying
function of the spectral index $n$. Thus, without even analyzing the
details of the spectral profile, explaining unknown fluxes via PBH
evaporation is far from compelling.
\section {The Initial Mass Function}
Carr \cite{carr} first calculated the PBH spectrum resulting from a
scale-invariant Harrison-Zel'dovich spectrum up to an overall
normalization. Subsequently Page and Hawking calculated a bound on the
normalization by calculating the expected diffuse gamma-ray spectrum
from these PBHs \cite{diffuse,mac2}. However, using the COBE measurements
of the temperature anisotropies translated into density fluctuations
(within a CDM), the overall normalization can be determined.
For a scale-invariant spectrum, no significant number density of PBHs
is generated. However, for a tilted blue power spectrum
with more power on smaller scales, a larger number density
of PBHs is expected.
Given an initially overdense region with density contrast $\delta_i$
and radius $R$ at time $t_i$ (using the usual comoving coordinates),
analytic arguments predict \cite{carrhawking,carr} a black hole will
form if
\begin{equation}
1/3\leq \delta_i \leq 1 \; .
\end{equation}
The lower bound on the density contrast comes from insisting that the
size of the region at the time of collapse be greater than the Jeans
length, while the upper bounds come from the consistency of the
initial data with the assumption of a connected topology.
The first calculations of the PBH mass distribution assumed the relation
\begin{equation}
\label{bhPropMh}
M_{bh}\simeq\gamma^{3/2}M_h,
\end{equation}
where $\gamma$ determines the equation of state $p = \gamma \rho$ and
$M_h$ is the horizon mass when the scale of interest crossed the
horizon. Recently, numerical evidence suggests
that near the threshold of black hole formation, gravitational collapse
behaves as a critical phenomena with scaling and self-similarity
\cite{scaling}. A scaling relation of the following form was found
\begin{equation}
\label{scaling}
M_{bh}(\delta)=kM_h(\delta-\delta_c)^{\rho},
\end{equation}
where $\rho$ is a universal scaling exponent which is independent
of the initial shape of the density fluctuation. It was
later shown \cite{JZI} that such scaling should be relevant for
PBH formation. Indeed, in Ref.~\cite{JZII} the authors found relation
(\ref{scaling}) to hold for PBH formation when the initial conditions
are adjusted to be nearly critical. They found the exponent to be
$\rho\approx 0.37$. They also found that for several different initial
density shapes, $\delta_c\approx 0.7$, which is significantly larger
than the analytic prediction of $1/3$ found by requiring that the
initial overdensity be larger than the Jeans mass.
Given Eq.~(\ref{scaling}), calculating the initial PBH mass
distribution becomes analytically cumbersome, since in principle one
needs to sum over all epochs of PBH formation. However, we would
expect that the initial mass function would be dominated by the
earliest epoch of formation if we assume a Gaussian distribution with
a blue power spectrum, since for larger scales the formation probability
should be suppressed. This expectation was tested in
Ref.~\cite{green2}, where the authors used the excursion set formalism
\cite{bond} to calculate the initial mass function allowing for PBH
formation at all epochs. The authors found that it was approximately true
the
that earliest epoch dominates, for the conditions
of interest to us.
This simplification allows us to derive quite easily the initial mass
distribution \cite{JZI}. We assume Gaussian fluctuations (the effects
of non-Gaussianity will be briefly discussed in Sec.~\ref{robust-sec})
and define the usual smoothed density contrast
\begin{equation}
\delta_R(x)=\int d^3y\,\delta (x+y) W_R(y) \; ,
\end{equation}
where $\delta(x)=(\rho(x)-\rho_b)/\rho_b$, and $\rho_b$ is background
energy density. $W_R$ is the window function with support in a region
of size $R$. The probability that a region of size $R$ has density
contrast between $\delta$ and $\delta+d\delta$ is given by
\begin{equation}
P(R,\delta)d\delta =\frac{1}{\sqrt{2 \pi}\sigma_R}\exp\left(-\frac{\delta^2}
{2\sigma_R^2}\right) d\delta \; ,
\end{equation}
where $\sigma_R$ is the mass variance for a region of size $R$,
$\sigma^2_R=\langle \delta_R^2(x) \rangle/R^3$.
Then using Eq.~(\ref{scaling}), the physical number density of PBHs
within the horizon per logarithmic mass interval, at the formation
epoch, can be written as
\begin{equation}
\label{IMF1}
\frac{d n_{bh}}{d\log M_{bh}} =
V_h^{-1}\,P[\delta(M_{bh})]\,\frac{d\delta}{d\log M_{bh}} =
\frac{V_h^{-1}}{\sqrt{2 \pi}\,\sigma\,\rho}
\left( \frac{M_{bh}}{k M_H} \right)^{1/\rho}\,
\exp\left[-\frac{1}{2\sigma^2} \left[ \delta_c
+ \left(\frac{M_{bh}}{k M_H} \right)^{1/\rho} \right]^2
\right],
\end{equation}
where $V_{h}$ is the horizon volume the the epoch of PBH formation.
We assume prompt reheating, and therefore take the formation epoch to
be the time of reheating, which corresponds to the minimum horizon
mass~\cite{green2}.
Assuming that the power law spectrum holds down to the
scales of the reheat temperature, then $\sigma_H^2 \propto R^{-(n+3)}$.
We can then relate the mass variance today $\sigma_0$ to the mass
variance at the epoch of PBH formation $\sigma(M_H)$, using
\cite{green,green2}
\begin{equation}
\label{dis1}
\sigma^2(M_H)=\sigma_0^2 \left( \frac{M_{eq}}{M_0} \right)^{(1-n)/3}
\left( \frac{M_{H}}{M_{eq}} \right)^{(1-n)/2},
\end{equation}
where
\begin{equation}
M_H = M_0\left(\frac{T_{eq}}{T_{RH}}\right)^2
\left(\frac{T_0}{T_{eq}}\right)^{3/2},
\end{equation}
$M_0$ is the mass inside the horizon today, $T_{RH}$ is the reheat
temperature, and $T_{eq}$ is the temperature at radiation-matter equality.
This relation is essential to connect present-day density fluctuations
to those of much earlier times. Two important assumptions underly
this useful relation: First, $n$ is taken to be constant over all
the scales of interest. Second, the universe was assumed to be
radiation dominated until the temperature dropped below
$T_{eq} \sim 5$~eV (matter-radiation equality), and then matter
dominated thereafter.
From the COBE anisotropy data, the mass variance can be calculated
\cite{bunn,green}
\begin{equation}
\label{COBE}
\sigma_0=9.5 \times 10^{-5} .
\end{equation}
Using this result we can then calculate the physical number density
per unit mass interval at $T = T_{RH}$
\begin{equation}
\label{IMF}
\frac{d n_{bh}}{dM_{bh}}=
\frac{V_h^{-1}}{\sqrt{2\pi}\sigma(M_H)M_{bh}\rho}y^{1/\rho}\exp\left[
-\frac{\left(\delta_c+y^{1/\rho}\right)^2}{2\sigma^2(M_H)}\right],
\label{phys-num-density-eq}
\end{equation}
where
\begin{equation}
y=\frac{M_{bh}}{kM_H}=\frac{M_{bh}T_{RH}^2}
{0.301 k g_\star^{-1/2} M_{Pl}^3} \; ,
\end{equation}
and $M_{Pl}$ is the Planck mass. The physical number density
at time $t$ is simply Eq.~(\ref{phys-num-density-eq}) rescaled
by a ratio of scale factors, that can be written as
\begin{equation}
\frac{d n_{bh}}{d M_{bh}}(t) = \left( \frac{T(t)}{T_{RH}} \right)^3
\frac{d n_{bh}}{d M_{bh}} \; .
\label{scaled-density-eq}
\end{equation}
Finally, we should note that relation (\ref{scaling}) is only valid
for $\delta \approx \delta_c$. Thus we expect that we may integrate
over $\delta$ with small errors, as long as the width of the Gaussian
is sufficiently small. In particular, our results should be trustworthy
provided $\sigma \lesssim 1$, which implies $n$ should not exceed
the maximum
\begin{equation}
n_{max}\simeq 1+\frac{2 \log(\sigma_0^2)}{\log(T_{eq}T_0/T_{RH}^2)} \; .
\end{equation}
We will see that $n$ does not exceed this maximum value for
all of the bounds we consider.
\section{Bounds from $\Omega \leq 1$}
Let us now calculate the total energy density in PBHs. We assume the
``standard cosmology'' where the universe began in an inflationary
phase, reheated, was radiation dominated from the reheating period
until matter-radiation equality, and then has been matter dominated.
The contribution of a PBH with a given initial mass, $M_{bh}$,
to the energy density today will depend upon its lifetime.
The time-dependent PBH mass $M(t)$ is given
by~\cite{hawking}
\begin{equation}
\label{massevol}
M(t) = M_\star\left[\left(\frac{M_{bh}}{M_\star}\right)^3
-\frac{t}{t_0}\right]^{1/3},
\end{equation}
where $M_\star$ is the initial mass of a PBH which would be decaying
today, $M_\star \simeq 5 \times 10^{14}$~g. It is a good approximation
to assume that the black hole decays instantaneously at a fixed decay
time, $t_d$, which we use in the following.
There are two components to the PBH density bounds that we can
calculate. The first is the total energy density of the PBHs that
have not decayed by a given time $t$. The second is the total
energy density of the products of PBH evaporation. The sum of these
components must be less than the critical density
$\Omega_{pbh,evap} + \Omega_{pbh} < 1$, at any time.
The evaporated products of PBHs, in particular photons,
could break up elements during nucleosynthesis, disrupting the
well-measured elemental abundances. This and other processes
during nucleosynthesis provide additional bounds on the density of
PBHs~\cite{nuc} that we do not discuss here.
The simplest bound comes from the density of PBHs that have not
decayed by time $t$,
\begin{equation}
\rho_{pbh}(t) = \left( \frac{T(t)}{T_{RH}} \right)^3
\rho_{tot}(t_{RH}) \> I_{M_\star(t)}^{M_{max}}(0)
\end{equation}
where $T(t)$ is the temperature of the universe at time $t$, and $M_\star(t)$
is the initial PBH mass that has just completely evaporated by time $t$,
\begin{equation}
M_\star(t) \approx M_\star \left( \frac{t}{t_0} \right)^{1/3} \; .
\end{equation}
$I_{M_1}^{M_2}(\xi)$ is a dimensionless weighted integral over the
PBH mass spectrum between $M_1$ to $M_2$,
normalized to the total density $\rho_{tot}(t_{RH})$,
\begin{equation}
I_{M_1}^{M_2}(\xi) = \frac{1}{\rho_{tot}(t_{RH})}
\int_{M_1}^{M_2} d M_{bh} M_{bh}
\frac{d n_{bh}}{d M_{bh}}
\left( \frac{M_{bh}}{M_\star} \right)^\xi \; ,
\end{equation}
where $d n_{bh}/d M_{bh}$ is given by Eq.~(\ref{phys-num-density-eq}).
We can use the above to trivially compute the ratio of the PBH density to
the critical density, $\Omega(t)$. In particular, we need
only compute the density ratio at three relevant epochs: immediately after
reheating $t=t_{RH}$, at matter-radiation equality $t=t_{eq}$, and
present-day $t=t_0$. The density ratios are\footnote{Note that
$\Omega(t_{RH}) = I_0^{M_{max}}(0)$ is often denoted by $\beta(t_{RH})$.}
\begin{eqnarray}
\Omega(t_{RH}) &=& I_0^{M_{max}}(0) \\
\Omega(t_{eq}) &=& \frac{T_{RH}}{T_{eq}}
I_{M_{eq}}^{M_{max}}(0) \\
\Omega(t_{0}) &=& \frac{T_{RH}}{T_{eq}}
I_{M_\star}^{M_{max}}(0) \; .
\end{eqnarray}
where $M_{max}$ corresponds to $\delta = 1$, which is of order the
horizon mass at reheating. The integral should be independent of the
upper limit $M_{max}$ if we are to trust our results.
Making the conservative approximation that all the PBH decay
products are relativistic, the contribution to the density ratio of
the products of PBH evaporation that has occurred up until today can
be written as
\begin{equation}
\Omega_{pbh,evap}(t_0) = \frac{T_{RH}}{T_{eq}} \left( \frac{T_0}{T_{eq}}
\right)^{1/4} I_{0}^{M_{eq}}(3/2) +
\frac{T_{RH}}{T_{eq}} I_{M_{eq}}^{M_\star}(2) \; .
\end{equation}
In Fig.~\ref{bound-fig}, we show the upper limit on $n$ as a function
of $T_{RH}$ coming from bounding $\Omega_{pbh,evap}(t_0) +
\Omega_{pbh}(t_0) < 1$ (solid line). For larger values of the reheat
temperature we get a more stringent bound by imposing the constraint
$\Omega_{pbh}(t_{RH}) \leq 1$, simply because as $T_{RH}$ is increased
more of the black holes will have decayed at an earlier epoch. Given
that most of the energy of the decay products resides in radiation,
the effect on $\Omega_{pbh}(t_0)$ is diminished due to the redshifting.
This new bound is given by the dotted line in Fig.~\ref{bound-fig}.
If we assume that the PBH leaves behind a Planck mass remnant, then we
have additional bounds which become important for very large reheat
temperature\cite{mac,barrow,cgl}. The best bound in this case comes
from calculating $\Omega_{remnant}(t_{eq})$ which is given by
\begin{equation}
\Omega_{remnant}(t_{eq}) = \frac{T_{RH}}{T_{eq}} \frac{M_{Pl}}{M_\star}
I_0^{M_{eq}}(-1) \; .
\end{equation}
The bound in this case is shown as the dashed line in Fig.~\ref{bound-fig},
and is the best bound at the largest values of the reheat temperature.
\section{Bounds from Diffuse Gamma-Rays}
For a certain range of $T_{RH}$ we can improve our bounds on $n$ from
diffuse gamma-ray constraints. The present day flux is determined by
convoluting the initial mass function with the black hole emission
spectrum
\begin{equation}
\label{hawkingspectrum}
f(x)=\frac1{2\pi}\frac{\Gamma_s(x)}
{\exp(8\pi x)-(-1)^{2 s}},
\end{equation}
where $s$ is the spin of the emitted particle, $x=\omega(t) M(t)/M_{Pl}^2$,
$\omega(t)$ is the frequency and $M(t)$ is the PBH mass at the time $t$
of emission. $\Gamma_s(x)$ is the absorption coefficient and may be
written as $[\omega(t)]^2 \sigma_s/\pi$.
$\sigma_s$ is the absorption cross section and is calculated using the
principle of detailed balance.
The values for $\sigma_s$ were calculated some time ago by
Page\cite{page1,page2}. Let us consider how $\sigma_s$
behaves for massless particles. At large values of $x$, $\sigma_s$
performs small oscillations about the geometric optics limit of
$\sigma_g=27 \pi M^2/M_{Pl}^4$. As $x$ approaches zero, $\sigma_s$
goes to zero for $s=1/2,1$ but goes to a constant value for $s=0$. We
will use the approximation
\begin{equation}
\Gamma_s(x)= (56.7,~20.4) x^2/\pi~~ $\rm for$~~ s=(\frac12,~1).
\end{equation}
This approximation is poor at low energies, as it is in
error by $50\%$ at $x=0.05$. However, as we shall see, the
contribution to the spectrum of interest is greatly peaked at
$x\simeq0.2$. The case of strongly interacting particles is
complicated by the hadronization process. There is a large
contribution coming from pion decay, however, given the extreme
sensitivity of the flux to the value $n$, the effect on the bound is
negligible.
It has been recently suggested \cite{heckler1,heckler2} that the
self-interactions of the emitted particles will induce a photosphere, thus
distorting the spectrum considerably from Eq.~(\ref{hawkingspectrum}).
It was suggested that two types of photospheres should form. A QCD
photosphere\footnote{In the case of QCD what is meant by
``photosphere'' is a quark-gluon cloud.} generated by parton-parton
interactions as well as a QED photosphere generated by
electron-positron-photon interactions. This is idea has been
tested more quantitatively via a numerical solution of the Boltzmann
equation~\cite{mcgill}.
Again, while this effect may change the spectrum, especially
at higher energies, it is irrelevant as far as the bound
on the spectral index is concerned.
The flux measured today is given by
\begin{equation}
\label{cosmflux}
\frac{dJ}{d\omega_0} = \frac{1}{4\pi} \int_{t_i}^{t_0} dt (1+z)
\int d M_{bh} \, \frac{d n_{bh}}{dM_{bh}}(t) \, f(x) \; ,
\end{equation}
where $d n_{bh}/d M_{bh}$ is evaluated at time $t$ using
Eq.~(\ref{scaled-density-eq}), $t_0$ is the age of the universe,
$t_i$ is the time of last scatter, and $f(x)$ is the instantaneous
emission spectrum given above with
\begin{equation}
x = \frac{\omega(t) M(t)}{M_{Pl}^2} = \frac{\omega_0 (1+z)}{M_{Pl}^2}
M_\star \left[ \left( \frac{M_{bh}}{M_\star} \right)^3
- \frac{t}{t_0} \right]^{1/3} .
\end{equation}
The integral over $t$ is cut off at early times, since at redshifts
above $z = z_0 \simeq 700$ the optical depth will be larger than unity
due to either pair production off of matter or ionized matter
\cite{ZS}. Those processes will degrade the energy below the window
we are interested in.
This integral may be rewritten in the more illuminating form
\begin{equation}
\label{cosmoSpectrum}
\frac{dJ}{d\omega_0} =
\frac{1}{4\pi}\frac{M_{Pl}^6}{(\omega_0M_\star)^3}
\int_{0}^{z_0}\frac{dz}{ H_0 (1+z)^{5/2}}
\int_{0}^{\infty} dx \> x^2 \alpha^{-2}\,f(x)\frac{d n_{bh}(x,z)}{dM_{bh}},
\end{equation}
where
\begin{equation}
\label{mbh}
\alpha=\frac{M(t)}{M_\star}=\left\{(1+z)^{-3/2} +
\frac{x^3M_{Pl}^6}{[(1+z)\omega_0M_\star]^3}\right\}^{1/3}.
\end{equation}
Let us study the qualitative behavior of the above integral as a
function of $\omega$ at fixed $n$ and $T_{RH}$. The $x$ integration is
controlled by the Boltzmann factor in $f(x)$. Indeed, a little
manipulation shows the the integrand is highly peaked near
$x \simeq 0.2$. Furthermore, the $\omega$ dependence in $\alpha^{-2}$
is almost completely canceled by the $\omega$ dependence in the
factor $M_{bh}^{-1}y^{1/\rho}\propto\alpha^{(1/\rho-1)}$ in
$d n_{bh}/dM_{bh}$. Thus the $\omega$ dependent part of the integrand may be
written as
\begin{equation}
\frac{dJ}{d\omega_0}\propto \omega_0^{-3}
\exp\left[-\frac{\left(\delta_c+aT_{RH}^{2/\rho}\alpha^{1/\rho}\right)^2}
{2\sigma^2(M_H)}\right],
\end{equation}
where $a^\rho = M_\star g_\star^{1/2}/(0.301 k M_{Pl}^3)$, and the
only $\omega$ dependence in the exponential is through $\alpha$. If
for now we assume that the dominant contribution the higher energy
photons comes from recent decays ($z\sim 0$), and most of the support
for the $x$ integral comes with $x \sim 0.2$, then $\alpha$ simplifies to
\begin{equation}
\alpha \approx \left[ 1 + \left( \frac{0.2 M_{Pl}^2}{w_0 M_\star} \right)^3
\right]^{1/3}.
\end{equation}
As $\omega_0$ gets larger than $0.2 M_{Pl}^2/M_\star \sim 100\ \rm{ MeV} $,
$\alpha$ becomes independent of $\omega_0$ and therefore the flux
behaves as $dJ/d\omega_0 \propto \omega_0^{-3}$.
For lower energies we can make the
approximation $\alpha\sim 0.2 M_{Pl}^2/(w_0M_\star)$, and we would
expect that at some point the $\omega$ dependence in the exponential
will begin to dominate such that the flux should begin to rapidly decrease
as we go to lower photon energies. The energy at which the flux turns
over is determined by the competition between the two terms $\delta_c$ and
$a T_{RH}^{2/\rho}\alpha^{1/\rho}$ in the exponential, which is set by
the reheat temperature. As we lower the reheat temperature the
position of the kink moves to lower energies. If the reheat
temperature is higher than $T_{RH}\sim 10^9\ \rm{ GeV} $, however, the peak
will stay around 100 MeV, since at these temperatures the second term
in the exponential will always dominate. Indeed, we expect the
position of the fall off to be near
\begin{equation}
\label{kink}
\omega_{\rm kink}\simeq \min\left(100{\ \rm{ MeV} }\,,\,
\frac{0.2g_\star^{1/2}\,T_{RH}^2}
{0.301\,k\,M_{Pl}\,\delta_c^{\rho}}\right).
\end{equation}
Figure~\ref{examples-fig} shows the flux for fixed $n$ for a few
different reheat
temperatures. The position of the kink is well tracked by
Eq.~(\ref{kink}). Note however that the flux does not fall off
exponentially at energies below the kink. This is because as we go to
lower energies we pick up more of a contribution from higher
redshifts.
It is interesting to contrast this behavior with the flux calculated
assuming that the mass of a PBH formed at a given epoch is proportional to
the horizon mass at the time of collapse. In Refs.~\cite{kim1,kim2} the
authors calculated an initial mass function following the
Press-Schecter formalism, summing over all epochs and assuming the
relation $M_{bh}\simeq\gamma^{3/2} M_H$ at each epoch. They found
\begin{equation}
\label{KLM}
\frac{d n_{bh}}{dM_{bh}} = \frac{n+3}{4} \sqrt{\frac{2}{\pi}} \gamma^{7/4}
\rho_i M^{1/2}_{H_i}M_{bh}^{-5/2}\sigma_H^{-1}
\exp\left(-\frac{\gamma^2}{2\sigma_H^2}\right),
\end{equation}
where $\rho_i$ and $M_{H_i}$ are the energy density and horizon mass
at $T_{RH}$ and
\begin{equation}
\sigma_H=\sigma_0\left(\frac{M_{bh}}{\gamma^{3/2}M_0}\right)^{(1-n)/4}.
\end{equation}
This result reduces to the initial mass function first computed by
Page \cite{carr} for the Harrison-Zel'dovich spectrum with $n=1$ and
$d n_{bh}/dM_{bh} \propto M_{bh}^{-5/2}$. The $\omega_0$ dependence
of this result arises only through $M(t)$ given by Eq.~(\ref{mbh}).
Using this initial mass distribution in Eq.~(\ref{cosmflux}), we
expect, as in the previous case, $dJ/d\omega_0 \propto \omega_0^{-3}$
for larger energies, and exponential decay into the lower energies
(which will again be mollified from photons descending from higher
redshifts). However, for this case the position of the kink will be
fixed at around 100 MeV, independent of the reheat temperature.
Let us compare the above prediction with the recent COMPTEL and EGRET
data. The EGRET collaboration found that the flux in the energy range
$30\ \rm{ MeV} -100\ \rm{ GeV} $ is well fit by the single power law \cite{EGRET}
\begin{equation}
\label{EGRET}
\frac{dJ}{d\omega_0}=7.32\pm 0.34\times 10^{-9}
\left(\frac{\omega_0}{451\ \rm{ MeV} }\right)^{-2.10\pm 0.03}~
({\rm cm^2~sec~sr~MeV})^{-1},
\end{equation}
while the COMPTEL data \cite{COMPTEL} in the range $0.8$--$30\ \rm{ MeV} $
can be fit \cite{kribs} to the power law
\begin{equation}
\frac{dJ}{d\omega_0}=6.40\times 10^{-3}
\left(\frac{\omega_0}{1\ \rm{ MeV} }\right)^{-2.38}~
({\rm cm^2~sec~sr~MeV})^{-1}.
\end{equation}
Below $0.8\ \rm{ MeV} $ there is large increase in the measured flux. Thus,
the best bounds are found by comparing the measured flux to predicted
flux at $\omega_{\rm kink}$ or at $0.8\ \rm{ MeV} $, whichever is
larger. Because of the rapid rise of the predicted spectrum relative
to the measured spectrum, a change in the kink position can change the
bound on $n$ on the order of $0.01$, which we consider within the
accuracy of our calculation. The bounds on $n$ from the diffuse
gamma-rays are specified by the dot-dashed line in
Fig.~\ref{bound-fig}. The bound terminates when all but the
exponential tail of the PBHs decay prior to a redshift of $700$, since
the optical depth at such early times exceeds unity, as discussed
above.
We may compare our results to those derived by Yokoyama
\cite{yokoyama},\footnote{After this work was completed, we
became aware of Ref.~\cite{greensusy} that also utilized the
critical collapse initial mass function to derive bounds on the
PBH mass density by requiring that LSPs (in supersymmetric models)
not be overproduced.} where the author placed bounds on mass fraction
of PBHs at $t_{RH}$, $\beta(t_{RH}) = \Omega(t_{RH})$, using the
initial mass function, Eq.~(\ref{IMF}).
He found that the bounds on $\beta$ did not differ
significantly from the previous bounds derived using the standard
initial mass functions, except for the bounds coming from diffuse
gamma-rays. In the latter case, applicable for horizon masses in the
range $M_H\geq 5\times 10^{14}$~g, he found more stringent
constraints. Our bounds on $n$, translated into bounds on $\beta$,
agree with his bounds coming from energy density constraints except
for the case of larger reheat temperature, since we included the proper
scaling of the energy density of photons emitted after PBH decay. Thus our
bounds on $\beta$ can differ by many orders of magnitude. Our bounds
coming from diffuse gamma-rays can also differ by orders of magnitude,
but in this case it is for a different reason. Yokoyama determined
his bound on $\beta$ by imposing the constraint on $\Omega_{pbh}(t_0)$
derived in Ref.~\cite{mac2}. However, when we change the initial mass
function we also change the diffuse gamma-ray spectrum significantly
in both shape and normalization, as discussed above. Thus, it is
inappropriate to directly take the bounds from Ref.~\cite{mac2} and
apply them to the case with the new initial mass function,
Eq.~(\ref{IMF}). We find that our bounds on $\beta$ from diffuse
gamma-rays are more stringent than those determined in
Ref.~\cite{yokoyama} in the range $M_H>5 \times 10^{15}$~g by several
orders of magnitude.
\section{Robustness of the Bounds}
\label{robust-sec}
Let us consider the robustness of the bounds. We might worry that the
bounds are highly sensitive to the choice of parameters given the
sharpness of the initial mass function. Indeed, in the case where it
is assumed that the PBH mass is given by Eq.~(\ref{bhPropMh}), the
bounds are $n$ are exceptionally sensitive to the exactness of this
relation. This is clear from the exponential factor in
Eq.~(\ref{KLM}). Given the initial mass function calculated
by Jedamzik and Niemeyer, Eq.~(\ref{IMF}),
we must check the sensitivity to the parameters $\delta_c,~k$ and
$\sigma_0$. In Ref.~\cite{JZII}, the authors tested the scaling relation
(\ref{scaling}) using several different initial shapes density
perturbations shapes. They found $(\delta_c=0.70, ~k=11.9)$,
$(\delta_c=0.67, ~k=2.85)$, $(\delta_c=0.71, ~k=2.39)$, for Gaussian,
Mexican Hat and fourth order polynomial fluctuations, respectively.
We varied the value of $\delta_c$ between $0.60-0.80$ and found that
the bounds changed by at most $0.01$. The sensitivity to the
variation being maximal at the smaller values of $T_{RH}$. Given that
the initial mass function is peaked at a number smaller than $k M_h$, the
sensitivity is increased at smaller $T_{RH}$ because $\sigma$ is an
decreasing function of $T_{RH}$. Variations in $k$ are equivalent to
a scaling in $T_{RH}$. Thus, varying $k$ by an order of magnitude has
essentially no effect on the bound. Lastly, let us consider the
sensitivity to the parameter $\sigma_0$. The value we used for
$\sigma_0$ in Eq.~(\ref{COBE}) was calculated in Ref.~\cite{green2}
using the result \cite{bunn}\footnote{It should be emphasized
that this result assumed the spectra
can be approximated as a power law over the range of $k$ that COBE
probes. We are then making the further assumption that $n$ is
constant down to the mass scales of relevance for PBHs.}
\begin{equation}
\label{fit}
\delta_0=1.91\times 10^{-5} \frac{\exp[1.01(1-n)]}{\sqrt{1+0.75 r}} \; ,
\end{equation}
where $r$ is a measure of the size of the tensor perturbations. The
$1\sigma$ observational error being $7\%$. The fit, Eq.~(\ref{fit}),
is good to within $1.5\%$ everywhere within the region $0.7 \leq n
\leq 1.3$ and $0\leq r \leq 2$. The authors of Ref.~\cite{bunn} quote
a $9\%$ uncertainty in Eq.~(\ref{fit}) at $1\sigma$, once
uncertainties in the systematics and variations in the cosmological
parameters are taken into account. The value in Eq.~(\ref{COBE}) was
determined ignoring tensor perturbations.
Given that $\sigma_0$ scales with $\delta_0$ we find that
varying $\sigma_0$ at the $2\sigma$ level has no effect on our bound
at the level of $0.01$. On the other hand, including some contribution
from tensor perturbation will weaken the bound. We found that taking
$r=2$ weakened the bound by $0.01-0.02$ throughout the range in the
reheat temperature.
We can also consider the effects of a non-vanishing $\Omega_\Lambda$.
Bunn et~al.~\cite{bunn} extended their results to this case and found
\begin{equation}
\label{labmda}
\delta_0|_{\Omega_\Lambda}=1.91\times 10^{-5} \frac{\exp[1.01(1-n)]}
{\sqrt{1+(0.75-0.13\,\Omega_\Lambda^2)r}}\Omega_0^{-0.80-0.05\log{\Omega_0}}
\, \left(1+0.18\,(n-1)-0.03\,r\,\Omega_\Lambda \right) \> .
\end{equation}
If $0 \le r \le 2$, we can express $\delta_0|_{\Omega_\Lambda}$
extracted assuming a nonzero cosmological constant to a very good
approximation by a scaling of $\delta_0$ extracted without a
cosmological constant
\begin{equation}
\delta_0|_{\Omega_\Lambda} \approx
\Omega_0^{-0.80-0.05\log{\Omega_0}} \> \delta_0 \; ,
\end{equation}
(where $\Omega_0 + \Omega_\Lambda = 1$)
and thus $\sigma_0$ also acquires a correction. Consequently,
the bound on $n$ is shifted by
\begin{equation}
\Delta n \equiv n - n|_{\Omega_\Lambda} =
\frac{2 \> (-0.8 - 0.05 \ln \Omega_0) \, \ln \Omega_0}{42.9
+ \ln (T_{RH}/10^8 \; \mathrm{GeV})} \; .
\end{equation}
In Fig.~\ref{delta-n-fig} we show the above correction as a function
of $T_{RH}$ for several choices of $\Omega_\Lambda$. If we take
$\Omega_\Lambda \approx 0.7$ as recent observational data suggests,
our bounds on $n$ strengthen by about $0.03-0.06$ for $T_{RH}$ between
$10^{16}-10^{3}$ GeV respectively, as shown in Fig.~\ref{bound-lambda-fig}.
We can also calculate bounds on the mass variance at reheating
\cite{liddle-grqc} which is essentially model-independent.
If the relation Eq.~(\ref{dis1}) is violated by, for example, a power
spectrum with a spectral index that depends on scale, then our previous
bounds on $n$ cannot be applied. However, given a inflationary
model one could in principle calculate the power spectrum,
normalize to the COBE data at our present epoch,
and then match onto the mass variance at reheating.
In Fig.~\ref{sigma-fig} we show the bounds on
the mass variance from both the density bounds as well as the
bounds from the diffuse gamma-ray observations. Notice that
the diffuse gamma-ray observation bounds on $\sigma(M_H)$
are a significant improvement over the density bounds in the
applicable range of reheating temperatures.
Finally, we must address the issue of non-Gaussianity. It has been
pointed out \cite{bullock} that skewness could very well be important
for PBH formation given that its effects are amplified in the tail of
the distribution $P[\delta]$, which contributes to PBH formation. In
general, the amount of non-Gaussianity expected is highly model
dependent. Bullock and Primack investigated several inflationary
models to study the amount of non-Gaussianity one would expect at
larger values of $\delta$. They calculated $P[\delta]$ for three toy
models, and found in one case no deviation from Gaussianity and in the
other two found a significant suppression in the probability of of
large perturbations. However, as was pointed out in Ref.~\cite{green},
while these effects can drastically effect the PBH mass fraction
$\beta$, we expect that, even in the most extreme case considered in
Ref.~\cite{bullock}, the effect on the bound on $n$ is only at the
level of $0.05$. For hybrid inflation, where the approximation that
$n$ is constant actually holds, the perturbations are in fact
Gaussian due to the linear dynamics of the inflaton field \cite{Yi}.
Therefore, these bounds should be applied to specific models, with the
roughness of the bound determined by the deviations away from
Gaussianity.
\section{Conclusions}
We have calculated the density of primordial black holes using the
the near critical collapse mass function that results in a spectrum
of PBH masses for a given horizon mass. The normalization of the
PBH mass spectrum was determined using the COBE anisotropy data
that allowed us to set bounds on the spectral index $n$ as a function
on the reheat temperature. We find that restricting the density of PBHs
to be less than the critical density corresponds to the restriction that
the spectral index $n$ be less than about $1.45$ to $1.2$, throughout
the range of reheating temperatures resulting after inflation,
$10^{3}$ to $10^{16}$ GeV respectively.
(The precise limits are shown in Fig.~\ref{bound-fig}.)
For a smaller range of reheating temperatures, between about
$10^{7}$ to $10^{10}$ GeV, significant PBH evaporation occurs
when the optical depth of the universe is less than one.
Hence, we found a slightly stronger bound on the spectral index
by restricting the cosmological PBH evaporation into photons
to be less than the present-day observed diffuse gamma-ray flux.
Due to the extreme sensitivity of the PBH mass density to the spectral index,
effects such as the indirect photon flux from PBH evaporation into
quarks and gluons which fragment into pions or the formation of a
QCD photosphere are completely negligible when calculating the bound
on $n$. We should also remark that slightly stronger bounds on $n$ for
larger reheating temperatures $\gtrsim 10^{10}$ GeV are expected
from PBHs that decay during the epoch of nucleosynthesis.
If the universe is vacuum-energy dominated, there are corrections
to our bounds on $n$ that can be substantial. We calculated these
corrections for a range of $\Omega_\Lambda$ and applied them to
our bounds on $n$ for the case of $\Omega_\Lambda = 0.7$. The
improvement is apparent by contrasting Fig.~\ref{bound-fig} with
Fig.~\ref{bound-lambda-fig}. Finally, we calculated bounds on the mass
variance at reheating. These bounds in principle could be used to
constrain any given inflationary model, once the power spectrum
is calculated.
\acknowledgments
This work was supported in part by the Department of Energy under
grant number DOE-ER-40682-143. We thank Rich Holman and Jane MacGibbon
for useful discussions. We also thank Andrew Liddle useful
discussions and comments on the manuscript.
{\tighten
|
\section{#1}} {\setcounter{equation}{0}}}
\newtheorem{th}{Theorem}[section]
\newtheorem{lm}{Lemma}[section]
\newtheorem{prop}{Proposition}[section]
\newtheorem{de}{Definition}[section]
\newtheorem{co}{Corollary}[section]
\newtheorem{re}{Remark}[section]
\newtheorem{con}{Conjecture}[section]
\def{K\"{a}hler }{{K\"{a}hler }}
\def{K\"{a}hler-Einstein }{{K\"{a}hler-Einstein }}
\def{Calabi-Yau }{{Calabi-Yau }}
\def{Lagrangian }{{Lagrangian }}
\def{special Lagrangian }{{special Lagrangian }}
\input epsf
\begin{document}
\hbadness=10000
\title{{\bf Lagrangian torus fibration of quintic {Calabi-Yau } hypersurfaces I:}\\
{\Large {\bf Fermat quintic case}}}
\author{Wei-Dong Ruan\\
Department of mathematics\\
Columbia University\\
New York, NY 10027\\
<EMAIL>}
\date{Revised December 1999}
\footnotetext{Partially supported by NSF Grant DMS-9703870.}
\maketitle
\begin{abstract}
In this paper we give a construction of Lagrangian torus fibration for Fermat type quintic {Calabi-Yau } hypersurfaces via the method of gradient flow. We also compute the monodromy of the expected {special Lagrangian } torus fibration and discuss structures of singular fibers.
\end{abstract}
\tableofcontents
\se{Introduction and background}
In this paper we give a construction of Lagrangian torus fibration for Fermat type quintic {Calabi-Yau } manifolds via the method of gradient flow. This method will produce Lagrangian torus fibration for {Calabi-Yau } hypersurfaces in general toric varieties. The results for general quintic hypersurfaces appeared in \cite{lag3}; for {Calabi-Yau } hypersurfaces in general toric varieties are written in \cite{tor}.\\\\
The motivation of our work comes from the study of Mirror Symmetry. Mirror Symmetry conjecture originated from physicists' work in conformal field theory and string theory. It proposes that for a {Calabi-Yau } 3-fold $M$ there exists a {Calabi-Yau } 3-fold $W$ as its mirror manifold. The quantum geometry of $M$ and $W$ are closely related. In particular one can compute the number of rational curves in $M$ by solving the Picard-Fuchs equation coming from variation of Hodge structure of $W$. \\\\
The first example of mirror symmetry was the Fermat type quintic given in the calculations by Candelas et al (\cite{Can}). Numerous examples of such calculations were worked out after that. A more general construction via toric varieties was given by Batyrev (\cite{Bat}). For a more complete history and reference the readers can consult \cite{mirrorbook}.\\\\
In 1996 Strominger, Yau and Zaslow (\cite{SYZ}) proposed a geometric construction of mirror manifold via special Lagrangian torus fibration. According to their program (we will call it SYZ construction), a {Calabi-Yau } 3-fold should admit a special Lagrangian torus fibration. The mirror manifold can be obtained by dualizing the fibers. Or equivalently, the mirror manifold of $M$ is the moduli space of special Lagrangian 3-torus in $M$ with a flat $U(1)$ connection. This conjectural construction was the first to give the mirror manifold directly from a {Calabi-Yau } manifold itself. \\\\
The notion of special Lagrangian submanifolds was first given by Harvey and Lawson in their celebrated paper \cite{HL}. For a {Calabi-Yau } manifold $(X, \Omega, \omega_g)$, where $\Omega$ is a holomorphic $(n,0)$ form and $\omega_g$ is the {K\"{a}hler } form of the {Calabi-Yau } metric, a {Lagrangian } submanifold $L$ in $X$ is called {special Lagrangian } if $L$ is area minimizing, or equivalently $\Omega$ restrict to $L$ is a constant multiple of the volume form of $L$.\\\\
According to the SYZ construction, {special Lagrangian } submanifold and {special Lagrangian } fibration for {Calabi-Yau } manifolds seem to play very important roles in understanding mirror symmetry. However, despite its great potential in solving the mirror symmetry conjecture, our understanding in special Lagrangian submanifolds is very limited. The known examples are mostly explicit local examples or examples coming from $n=2$. There are very few examples of {special Lagrangian } submanifold or {special Lagrangian } fibration for dimension higher than two. M. Gross, P.M.H. Wilson, N. Hitchin, P. Lu and R. Bryant(\cite{Gross1}\cite{Gross2}\cite{GW}\cite{H}) did some work in this area in recent years. On the other extreme, in \cite{Z}, Zharkov constructed some non-{Lagrangian } torus fibration of {Calabi-Yau } hypersurfaces in toric variety.\\\\
When dimension $n=2$, {Calabi-Yau } manifold is hyper{K\"{a}hler }. Therefore for any {Calabi-Yau } metric, there is an $S^2$ family of complex structures that are compatible with the given {Calabi-Yau } metric. Special Lagrangian submanifold for one compatible complex structure exactly corresponds to complex curves for another compatible complex structure. Therefore {special Lagrangian } theory is reduced to the theory of complex curves in $X$, which is fairly well understood. For $n\geq 3$, there are no nice interpretations like this. \\\\
Given our lack of knowledge for {special Lagrangian }, one may consider relaxing the requirement to consider {Lagrangian } fibration. Special Lagrangians are very rigid and hard to find. On the other hand, {Lagrangian } submanifolds are more flexible and can be modified locally by Hamiltonian deformation. This is a reasonable first step to take. For many applications to mirror symmetry, especially those concerning (symplectic) topological structure of fibration, {Lagrangian } fibration will provide quite sufficient information. In this paper we mainly consider Lagrangian torus fibration of {Calabi-Yau } hypersurfaces in toric varieties. \\\\
Our idea is a very natural one. We try to use gradient flow to get Lagrangian torus fibration from a known Lagrangian torus fibration at the "Large Complex Limit". It will in principle be able to produce {Lagrangian } fibration in general {Calabi-Yau } hypersurfaces. For simplicity, it is helpful to explore the case of Fermat type quintic {Calabi-Yau } threefold in ${\bf CP^4}$ in detail first, which is the case studied in \cite{Can}. Most of the essential features for {Calabi-Yau } hypersurfaces in general toric varieties already show up here.\\\\
The paper is organized as follows: In Section 2 we will first describe a {Lagrangian } torus fibration for Fermat type quintic {Calabi-Yau } familly $\{X_\psi\}$ in ${\bf P^4}$\\
\[
p_{\psi}=\sum_1^5 z_k^5 - 5\psi \prod_{k=1}^5 z_k=0
\]\\
at the large complex limit $X_{\infty}$\\
\[
p_{\infty}=\prod_{k=1}^5 z_k=0.
\]\\
In Section 3 we construct an appropriate vector field whose gradient flow will give {Lagrangian } fibration to nearby {Calabi-Yau } hypersurfaces. In Section 4, we will discuss the expected fibration structure by explicitly computing the monodromy transformations of the expected fibration if the fibration is actually {special Lagrangian } fibration. In Section 5 we will describe the expected structures of singular fibers implied by monodromy information. In Section 6 we will compare our {Lagrangian } fibration with the expected {special Lagrangian } fibration---their topological structure are not exactly the same; we will discuss the cause for their differences. Finally in Section 7, we will discuss the relevance to mirror construction for toric {Calabi-Yau } manifolds through the dual polyhedra construction.\\\\
This paper is the first part of several papers in this subject. In \cite{lag3} we discuss general construction of Lagrangian torus fibration for general quintic (non-Fermat) {Calabi-Yau } hypersurfaces. In \cite{lag2} we address the technical problem of modifying our construction into a {Lagrangian } torus fibration with the topological type of the expected one. \\\\
\se{{Lagrangian } torus fibration for large complex limit}
Consider the well studied Fermat type quintic {Calabi-Yau } familly $\{X_\psi\}$ in ${\bf P^4}$ defined by\\
\[
p_{\psi}=\sum_1^5 z_k^5 - 5\psi \prod_{k=1}^5 z_k=0.
\]\\
When $\psi$ approaches $\infty$, The familly approach its ``large complex limit'' $X_{\infty}$\\
\[
p_{\infty}=\prod_{k=1}^5 z_k=0.
\]\\
$X_{\infty}$ is a union of five ${\bf P^3}$. There is a natural degenerate $T^3$ fibration structure for $X_{\infty}$. Let $\{P_i\}_{i=1}^5$ be five points in ${\bf R^4}$ that are in general position. Consider the natural map $F: \ P^4\longrightarrow R^4$.\\
\[
F([z])=\sum_{k=1}^5 \frac{|z_k|^2}{\sum_{i=1}^5 |z_i|^2}P_k.
\]\\
Then $\Delta={\rm Image}(F)$ is a 4-simplex. $X_{\infty}$ is naturally fibered over $\partial \Delta$ via this map $F$ with general fiber being $T^3$. This is precisely the $T^3$ special Lagrangian fibration for $X_{\infty}$ as indicated by SYZ construction.\\\\
To see this fact, we need to construct a suitable {Calabi-Yau } metric on $X_{\infty}$ such that $F$ is a {special Lagrangian } torus fibration with respect to the {K\"{a}hler } form of the {Calabi-Yau } metric. Since $X_{\infty}$ is just a union of several ${\bf CP^3}$'s, we will concentrate on one of these ${\bf CP^3}$. We will actually carry out the discussion for general ${\bf CP^n}$.\\\\
In general, on $ {\bf C^{n+1}} $ there is a natural flat {K\"{a}hler } metric with {K\"{a}hler } form\\
\[
\omega_0 = \sum_{i=1}^{n+1} \frac{dz_i\wedge d\bar{z}_i}{|z_i|^2}
\]\\
which is singular along the coordinate hyperplanes. Consider the natural projection $\pi: \ {\bf C^{n+1}}-\{0\}\rightarrow {\bf CP^{n}} $. The restriction of $\omega_0$ to the hypersurface $\prod_{i=1}^{n+1}z_i =1$ in $\bf C^{n+1}$ naturally push forward via $\pi$ to a flat {K\"{a}hler } metric $\omega$ on ${\bf CP^{n}}$ that is singular along the union of projective coordinate hyperplanes, which is exactly our large complex limit $Y_{\infty}$. (Here we use $Y_{\infty}$ to denote dimention $n-1$ version of the large complex limit in ${\bf CP^{n}}$ to distinguish from $X_\infty$ that corresponding to $n=4$ case.) Take local coordinate $x_i=\frac{z_i}{z_{n+1}}$ for $i=1,\cdots, n$. Then\\
\[
\omega = \sum_{i=1}^{n} \frac{dx_i\wedge d\bar{x}_i}{|x_i|^2} - \frac{1}{n+1} \left(\sum_{i=1}^{n} \frac{dx_i}{x_i}\right)\left(\sum_{i=1}^{n} \frac{d\bar{x}_i}{\bar{x}_i}\right)
\]\\
It is nice to compute their {K\"{a}hler } potentials:\\
\[
\omega_0 = \partial \bar{\partial} \left( \sum_{i=1}^{n+1} (\log|z_i|^2)^2 \right)
\]\\
\[
\omega = \partial \bar{\partial} \left( \sum_{i=1}^{n} (\log|x_i|^2)^2 - \frac{1}{n+1} (\sum_{i=1}^{n}\log|x_i|^2)^2\right)
\]\\
Clearly\\
\[
\omega^n = (-1)^{\frac{n(n-1)}{2}} \frac{n!}{n+1}\Omega \wedge \bar{\Omega}
\]\\
where\\
\[
\Omega = \bigwedge_{i=1}^n \frac{dx_i}{x_i}
\]\\
is the holomorphic n-form on ${\bf CP^{n}}\backslash Y_{\infty}$. Notice that $\omega$ corresponds to a complete {Calabi-Yau } metric on ${\bf CP^{n}}\backslash Y_{\infty}$. Consider the n-dimension real torus $T^n$ represented by\\
\[
e^{i\theta} = (e^{i\theta_1}, e^{i\theta_2}, \cdots, e^{i\theta_n})
\]\\
as a real abelian Lie group. Then $T^n$ act naturally on ${\bf CP^{n}}$\\
\[
e^{i\theta}(x) = (e^{i\theta_1}x_1, e^{i\theta_2}x_2, \cdots, e^{i\theta_n}x_n)
\]\\
as symplectomorphisms. The moment map is exactly $F: {\bf CP^{n}}\backslash Y_{\infty} \rightarrow {\bf R^n}$\\
\[
F(x)= ( \log |x_1|^2, \log |x_2|^2, \cdots, \log |x_n|^2)
\]\\
It is easy to see that when $\Omega$ is restricted to the fibre, we have\\
\[
\Omega|_{T^n(x)} = i^n\bigwedge_{i=1}^n d\theta_i
\]\\
Therefore $F$ naturally gives us the {special Lagrangian } fibration of ${\bf CP^{n}}\backslash Y_{\infty}$. Specialize to $n=3$ case, this construction gives us the {special Lagrangian } torus fibration with respect to a complete (flat) {Calabi-Yau } metric on $X_\infty\backslash {\rm Sing}(X_\infty)$.\\\\
When $\psi$ is large, according to SYZ conjecture, we {\it expect} that $X_{\psi}$ will also possess a special largrangian $T^3$ fibration with base identified with $\partial \Delta$ which is topologically an $S^3$. There are still serious analysis and geometric works to be done to totally justify the special Lagrangian fibration. We will instead give a natural construction of Lagrangian torus fibration.\\\\
For our purpose, we will consider the Fubini-Study metric\\
\[
\omega_{FS} = \partial \bar{\partial} \log(1+|x|^2).
\]\\
The nice thing about the Fubini-Study metric is that the $T^n$ action is also a symplectomorphism with respect to $\omega_{FS}$. The corresponding moment map is\\
\[
F(x)= \left( \frac{|x_1|^2}{1+|x|^2}, \frac{|x_2|^2}{1+|x|^2}, \cdots, \frac{|x_n|^2}{1+|x|^2}\right)
\]\\
which is easy to see if we write $\omega_{FS}$ in polar coordinates.\\
\[
\omega_{FS} = \partial \bar{\partial} \log(1+|x|^2)=i\sum_{j=1}^n d\theta_i \wedge d\left(\frac{|x_i|^2}{1+|x|^2}\right).
\]\\
The most symmetric expression of the {K\"{a}hler } potential for the Fubini-Study metric is\\
\[
h= \frac{1 + |x|^2}{\left(\prod_{i=1}^n|x_i|^2\right)^{\frac{1}{n+1}}} = \frac{|z|^2}{\left( \prod_{i=1}^{n+1}|z_i|^2 \right)^{\frac{1}{n+1}}}
\]\\
This expression can also be derived by restricting $|z|^2$ to the hypersurface $\prod_{i=1}^{n+1}z_i=1$ and then push down by $\pi$.\\\\
In general, we have the following lemma:\\
\begin{lm}
For a {K\"{a}hler } metric\\
\[
\omega = \partial \bar{\partial} h(x)
\]\\
$T^n$ acts as symplectomorphism if and only if $h$ can be chosen to depend only on $|x_i|^2$.
\end{lm}
{\bf Proof:} Assume that this is the case, then\\
\[
\omega = \partial \bar{\partial} h = i\sum_{j=1}^n d\theta_i \wedge d h_i.
\]\\
where\\
\[
h_i=|x_i|^2\frac{\partial h}{\partial |x_i|^2}.
\]\\
From above expression of $\omega$, clearly $T^n$ acts as symplectomorphism and\\
\[
F_h(x) = (h_1, h_2, \cdots, h_n)
\]\\
is the moment map.
\begin{flushright} $\Box$ \end{flushright}
For the purpose of this paper, any $T^n$-invariant {K\"{a}hler } metric as in Lemma 2.1 is as good. We will mainly use the Fubini-Study metric.\\\\
\se{Lagrangian torus fibration via a gradient flow}
\subsection{The gradient vector field}
Consider the meromorphic function\\
\[
s= \frac{\prod_{k=1}^5 z_k}{\sum_1^5 z_k^5}
\]\\
defined on ${\bf P^4}$. Let $\omega$ denote the {K\"{a}hler } form of a {K\"{a}hler } metric $g$ on ${\bf P^4}$, and $\nabla f$ denote the gradient vector field of real function $f=Re(s)$ with respect to the {K\"{a}hler } metric $g$. To describe the construction, we need the following facts.\\
\begin{lm}
The gradient flow of $f$ leaves the set $\{Im(s)=0\}$ invariant and deforms {Lagrangian } submanifolds in $X_{\infty}$ to {Lagrangian } submanifolds in $X_{\psi}$.\\
\end{lm}
{\bf Proof:} Clearly $\nabla f$ is always perpendicular to level sets of $f$. Therefore,\\
\[
J\nabla f (f)=0.
\]\\
Let $s = f+ ih$. Since $s$ is holomorphic, we have\\
\[
\nabla f (h) = - J\nabla f (f)=0.
\]\\
Therefore $h$ is constant along the gradient flow, or in another word, $\{Im(s)={\rm constant}\}$ is invariant under the gradient flow of $f$. In particular, $\{Im(s)=0\}$ is invariant under the gradient flow of $f$.\\\\
Notice the fact that\\
\[
\{Im(s)=0\} = \bigcup_{\psi ^{-1}\in {\bf R}} X_{\psi}
\]\\
Let $L$ be a {Lagrangian } submanifold of $X_{\infty}$. To prove the lemma, it is sufficient to show that the 4-dimension submanifold $S$ swept out by $L$ under the gradient flow of $f$ is {Lagrangian } submanifold of ${\bf P^4}$. \\\\
First, for a vector field $v$ on $L$, since $L\subset X_{\infty}$ and $\nabla f$ is perpendicular to level sets of $f$ (in particular, perpendicular to $X_{\infty}$), we see that $Jv$ is along $X_{\infty}$ and $\nabla f$ is perpendicular to $Jv$ along $L$. Therefore, $\omega(v,\nabla f)=0$. \\\\
Let $V$, $W$ denote vector fields on $S$ that are invariant under the gradient flow $\phi_t$. $\nabla f$ is a vector field of this type. Using the fact that $f$ is pluri-harmonic and $\omega_g$ is {K\"{a}hler } form, it is easy to derive that\\
\begin{eqnarray*}
\frac{d}{dt}\omega_g(V,W) &=& -{\cal L}_{\nabla f} \omega_g(V,W) = -(d i(\nabla f)\omega_g)(V,W)\\
&=& (dJdf)(V,W) = -2i\partial \bar{\partial} f(V,W)=0.
\end{eqnarray*}
Therefore $\omega_g(V,W)$ is constant along the flow. Since initial value of invariant vector fields are spanned by vector fields along $L$ and $\nabla f$, by the {Lagrangian } property of $L$ and the fact that $\omega(v,\nabla f)=0$ for vector field $v$ on $L$, we have $\omega_g(V,W)=0$, therefore $S$ is {Lagrangian }.
\begin{flushright} $\Box$ \end{flushright}
{\bf Remark:}\\
(i) The lemma can also be understood roughly by the fact: $\nabla f = H_h$ (the hamiltonian vector field generated by $h=Im(s)$). At the smooth part of the vector field, this fact essentially implies the lemma, although at singular part of the vector field, additional argument as in the proof is needed.\\\\
(ii) The proof of the lemma actually implies something more. Lower dimension torus in $X_{\infty}$ form critical set of $f$. The proof of the lemma also implies that the stable manifold of a lower dimensional torus with respect to flow of $\nabla f$ is {Lagrangian } in ${\bf P^4}$, therefore intersect with $X_{\psi}$ at a {Lagrangian } submanifold.
\begin{flushright} $\Box$ \end{flushright}
With this lemma in mind, the construction of {Lagrangian } torus fibration of $X_{\psi}$ for $\psi$ large is immediate. Recall that in last section we had a canonical Lagrangian torus fibration of $X_{\infty}$ over $\partial \Delta$. Deform along gradient flow of $f$ will naturally induce a {Lagrangian } torus fibration of $X_{\psi}$ over $\partial \Delta$ for $\psi$ large and real.\\\\
However $X_{\infty}$ is singular and $\nabla f$ is also singular where $s$ is singular. To get a really honest {Lagrangian } torus fibration, we need to discuss how to deal with these singularities.\\\\
Along the gradient flow of $f$ we have\\
\[
\frac{df}{dt} = \nabla f \cdot \frac{dx}{dt} = \nabla f \cdot \nabla f = |\nabla f|^2
\]\\
We see the gradient flow of $f$ does not exactly move $X_{\infty}$ to $X_{\psi}$. To ensure this property, the flow has to satisfy $\frac{df}{dt} = {\rm constant}$. This will be true if we scale the vector field $\nabla f$ to\\
\[
V=\frac{\nabla f}{|\nabla f|^2}.
\]\\
\subsection{Flow from the smooth part of $X_{\infty}$}
To understand the flow of $\nabla f$ and $V$, it is helpful to express them in local coordinate. Recall that $X_{\infty}=\cup_{k=1}^5 D_k$, where $D_k=\{z_k=0\}$. Let's choose coordinate $x_i=z_i/z_5$, for $i=1,2,3,4$ and consider $D_4$, where $x_4=0$. Under this coordinate, we have\\
\[
s=\frac{\prod_{i=1}^4x_i}{\sum_{i=1}^4x_i^5 +1}
\]\\
\[
ds =\frac{\prod_{i=1}^4x_i}{\left(\sum_{i=1}^4x_i^5 +1\right)^2}\sum_{i=1}^4 \left(\sum_{j=1}^4x_j^5 +1-5x^5_i\right)\frac{dx_i}{x_i}
\]\\
\[
ds|_{D_4} =\frac{\prod_{i=1}^3x_i}{\sum_{i=1}^3x_i^5 +1}dx_4
\]\\
For simplicity, we choose {K\"{a}hler } metric to be\\
\[
g=\sum_{i=1}^4 dx_i\wedge d\bar{x}_i
\]\\
Then when restricted to $D_4$, we have\\
\[
|ds|^2 = \left|\frac{\prod_{i=1}^3x_i}{\sum_{i=1}^3x_i^5 +1}\right|^2
\]\\
\[
\nabla s = \frac{\prod_{i=1}^3x_i}{\sum_{i=1}^3x_i^5 +1} \frac{\partial}{\partial \bar{x}_4}
\]\\
These give us\\
\[
V=\frac{\nabla f}{|\nabla f|^2} = Re\left(\frac{\sum_{i=1}^3x_i^5 +1}{\prod_{i=1}^3x_i} \frac{\partial}{\partial x_4}\right)
\]\\
So we see that {\bf the flow of $V$ is smooth when restricted to the smooth part of $X_{\infty}$.}\\\\
Another way to understand $V$ is to realize that $ds$ is a holomorphic section of $N^*_X$, $(ds)^{-1}$ is a natural holomorphic section of $N_X$. Notice the exact sequence\\
\[
0 \rightarrow T_X \rightarrow T_{\bf P^4}|_X \rightarrow N_X \rightarrow 0
\]\\
With respect to the {K\"{a}hler } metric $g$ on $T_{\bf P^4}$, the exact sequence has a natural (non-holomorphic) splitting\\
\[
T_{\bf P^4}|_X =T_X \oplus N_X
\]\\
$V$ is just real part of the natural lift of $(ds)^{-1}$ via this splitting. $V$ is singular exactly when $(ds)^{-1}$ is singular or equivalently, when $ds=0$, which corresponds to singular part of $X_{\infty}$. On the other hand, the union of the smooth three-dimensional {Lagrangian } torus fibers of $X_{\infty}$ is exactly the smooth part of $X_{\infty}$. All these 3-torus fibers will be carried to $X_{\psi}$ nicely by the flow of $V$.\\
\subsection{Flow from the singular part of $X_{\infty}$}
Now we will try to understand how the gradient flow of $f$ behaves at singularities of $X_{\infty}$. For this purpose, it is helpful to understand the following example.\\\\
{\bf Example:}
Consider holomorphic function $s(z)=e^{i\theta}z_1z_2\cdots z_n$ on ${\bf C^n}$. $X=\{s=0\}$ is a variety with normal crossing singularities. There is a natural map \\
\[
F:\ {\bf C^n} \rightarrow {\bf R^n_+},\ \ F(z_1,\cdots,z_n) = (|z_1|,\cdots,|z_n|).
\]\\
For $c\in {\bf R^n_+}$, the fiber\\
\[
F^{-1}(c) = \{z:|z_i|=c_i,\ {\rm for}\ i=1,\cdots,n \}
\]\\
is $n$-torus for generic $c$. Let $f=Re(s)$, then with respect to the flat metric on ${\bf C^n}$, we have\\
\[
\nabla f = {\rm Re}\left( s\sum_{i=1}^n \frac{1}{z_i}\frac{\partial}{\partial \bar{z}_i}\right)
\]\\
By our previous argument, $h(z) ={\rm Im}(s)=Im(e^{i\theta}z_1z_2\cdots z_n)$ is invariant under gradient flow of $f$. It is also easy to observe that $\nabla f$ will leave $h_{ij}(z)=|z_i|^2 -|z_j|^2$ invariant. For $c=(c_1,\cdots, c_n)$ such that $\sum c_i =0$, let\\
\[
S_c = \{z: h(z)=0,\ h_{ij} =c_{ij}=c_i-c_j,\ {\rm for}\ i,j=1,\cdots,n \}.
\]\\
It is easy to see that $S_c$ are special Lagrangians in ${\bf C^n}$. Actually this is an example mentioned in Harvey and Lawson's paper. Let $X_r = \{z: s(z)=r\}$, then $L_{c,r} = S_c\cap X_r$ give us a smooth {Lagrangian } $n$-torus fibration of $X_r$ for $r$ real.\\\\
Define\\
\[
\theta_{ij}= \theta_i-\theta_j = \frac{i}{2}\left( \log\frac{\bar{z}_i}{z_i} -\log\frac{\bar{z}_j}{z_j}\right),
\]\\
where $\theta_i$ is the argument of $z_i$. Then\\
\[
\nabla f (\theta_{ij}) =-\frac{1}{2} {\rm Im} \left(\frac{s}{|z_i|^2} - \frac{s}{|z_j|^2}\right)=-\frac{1}{2} \left(\frac{1}{|z_i|^2} - \frac{1}{|z_j|^2}\right)h(z).
\]\\
Recall that the real hypersurface $\{h(z)=0\}$ is invariant under the flow. Above equation implies that when restricted to $\{h(z)=0\}$ $\theta_{ij}$ is invariant. For $c=(c_1,\cdots, c_n)$ such that $\sum c_i =0$, let\\
\[
T_c = \{z: h(z)=0,\ \theta_{ij} =c_{ij}=c_i-c_j,\ {\rm for}\ i,j =1,\cdots,n \}.
\]\\
There is another very illustrative way to write $T_c$.\\
\[
T_c = \{z: \theta_i-\theta_j=c_{ij}=c_i-c_j, \theta+\sum_{i=1}^n\theta_i = 0\}=\{z:\theta_i=c_i-\frac{\theta}{n}\}
\]\\
It is easy to see that $T_c$ are special Lagrangians in ${\bf C^n}$ that is invariant under the flow. $T_{c,r} = T_c\cap X_r$ give us a smooth {Lagrangian } fibration on the horizontal directions of $X_r$ for $r$ real.
\begin{flushright} $\Box$ \end{flushright}
As we know, $X_{\infty}$ has only normal crossing singularities. The above example gave us a rough picture of the local behavior of the gradient flow of $f$ around singularities of $X_{\infty}$ {\bf when denominator of $s$ is non-zero}. For detailed discussion and proof, please refer to \cite{lag2}.\\\\
Finally it remains to analyze the case when denominator of $s$ is zero and $X_{\infty}$ is singular, which is the following set $\Sigma$. Let\\
\[
\Sigma_{ijk} =\{[z]\in {\bf CP^4}| z_i^5 + z_j^5 + z_k^5 =0, z_l=0\ \ {\rm for}\ l\in\{1,2,3,4,5\}\backslash\{i,j,k\}\}.
\]
$\Sigma_{ijk}$ is a genus 6 curve. Let\\
\[
\Sigma = \bigcup_{\{i,j,k\}\in \{1,2,3,4,5\}} \Sigma_{ijk}
\]
We see that $\Sigma = {\rm Sing}(X_{\infty})\cap X_{\psi}$ for any $\psi$.\\\\
In general, points in $X_{\infty}\cap X_{\psi}$ are fixed under the flow. In particular, $\Sigma \subset X_{\infty}\cap X_{\psi}$ is fixed under the flow. Since the vector field $V$ is singular along $\Sigma$, more argument is needed to ensure the flow behave as expected near $\Sigma$. For the argument to work, it is actually necessary to deform the {K\"{a}hler } form slightly near $\Sigma$. For details, please also refer to \cite{lag2}.\\
\subsection{Lagrangian torus fibration structure}
To understand the {Lagrangian } torus fibration of $X_{\psi}$, it is helpful to first review the torus fibration of $X_{\infty}$. For any subset $I \subset \{1,2,3,4,5\}$, Let\\
\[
D_I = \{z:\ z_i=0,z_j\not= 0,\ {\rm for}\ i\in I,j\in\{1,2,3,4,5\}\backslash I\},
\]\\
$\Delta_I = F(D_I)\subset \partial\Delta$. Let $|I|$ denote the cardinality of $I$. We have\\
\[
\partial\Delta = \bigcup_{\tiny{\begin{array}{c}I \subset \{1,2,3,4,5\}\\ 0<|I|<5\end{array}}} \Delta_I
\]\\
The fibers over $\Delta_I$ are $4-|I|$ dimensional torus.\\\\
The flow of $V$ moves $X_{\infty}$ to $X_{\psi}$. When $|I|=1$, the flow is diffeomorphism on $D_I$ and moves the corresponding smooth 3-torus fibration in $D_I$ to 3-torus fibration in $X_{\psi}$.\\\\
When $|I|>1$, by above discussion after the example, each point of $D_I\backslash \Sigma$ will be deform to a $|I|-1$ torus under the flow of $V$. Therefore a torus fiber in $D_I$ not intersecting $\Sigma$ will be deformed to a smooth 3-torus {Lagrangian } fiber of $X_{\psi}$ under the flow of $V$.\\\\
Now let's try to understand the singular fibers. Notice that $D_I\cap\Sigma\not= \emptyset$ if and only if $|I|=2,3$. Let $\tilde{\Gamma}_{ijk} = F(\Sigma_{ijk})$. There are 3 types of singular fiber over different portion of\\
\[
\tilde{\Gamma} = \bigcup_{\{i,j,k\}\in \{1,2,3,4,5\}} \tilde{\Gamma}_{ijk}
\]\\
Let $\tilde{\Gamma}^2$ denote the interior of $\tilde{\Gamma}$, and\\
\[
\tilde{\Gamma}^1 = \partial \tilde{\Gamma} \cap \left(\bigcup_{|I|=2}\Delta_I\right),
\]
\[
\tilde{\Gamma}^0 = \partial \tilde{\Gamma} \cap \left(\bigcup_{|I|=3}\Delta_I\right).
\]\\
Then\\
\[
\tilde{\Gamma} = \tilde{\Gamma}^0 \cup \tilde{\Gamma}^1 \cup \tilde{\Gamma}^2.
\]\\
Without loss of generality, let us concentrate on\\
\[
\overline{D}_{45} = \{z: z_4=z_5=0\}
\]\\
with the natural coordinate $x_i=z_i/z_3$ for $i=1,2$. Under this coordinate\\
\[
\Sigma_{123} =\{x=(x_1,x_2)| x_1^5 + x_2^5 + 1 =0\}.
\]\\
$r=(r_1,r_2)=(|x_1|,|x_2|)$ can be thought of as coordinate on $\overline{\Delta}_{45}$. Under this coordinate\\
\[
\tilde{\Gamma}_{123} = \{r=(r_1,r_2)|r_1^5 + r_2^5 \geq 1, r_1^5 +1 \geq r_2^5, r_2^5 +1 \geq r_1^5\}
\]\\
For simplicity, we will omit the index and denote $\Sigma_{123}$ by $\Sigma$ and $\tilde{\Gamma}_{123}$ by $\tilde{\Gamma}$. Then\\
\[
\tilde{\Gamma}^2 = \{r=(r_1,r_2)|r_1^5 + r_2^5 > 1, r_1^5 +1 > r_2^5, r_2^5 +1 > r_1^5 \},
\]
\[
\tilde{\Gamma}^0 = \{(1,0),(0,1)\},
\]
\[
\tilde{\Gamma}^1 = \{r=(r_1,r_2)|r_1^5 + r_2^5 = 1\ {\rm or}\ r_1^5 = r_2^5 + 1\ {\rm or}\ r_2^5 = r_1^5 + 1\}\backslash \tilde{\Gamma}^0.
\]\\
$(x_1,x_2)\rightarrow (e^{\frac{2\pi i}{5}k_1}x_1,e^{\frac{2\pi i}{5}k_2}x_2)$ give us a natural action of ${\bf Z_5\times Z_5}$ on $\Sigma$. Combine with the ${\bf Z_2}$ action $(x_1,x_2)\rightarrow (\bar{x}_1,\bar{x}_2)$ we have\\
\[
\begin{array}{ccc}
{\bf Z_2\times Z_5\times Z_5}&\rightarrow& \Sigma\\
&&\\
&&\downarrow^F\\
&&\\
&&\tilde{\Gamma}\\
\end{array}
\]\\
$F$ is a $50$-sheet covering map over $ \tilde{\Gamma}^2$. ${\bf Z_2}$ is the stablizer of the group action on $F^{-1}(\tilde{\Gamma}^1)\cap\Sigma$. For $p\in \tilde{\Gamma}^1$, $F^{-1}(p)$ is a 2-torus that intersect $\Sigma$ at $25$ points. ${\bf Z_2\times Z_5}$ is the stablizer of the group action on $F^{-1}(\tilde{\Gamma}^0)\cap\Sigma$. For $p\in \tilde{\Gamma}^0$, $F^{-1}(p)$ is a circle that intersect $\Sigma$ at $5$ points.\\\\
From our previous discussion, when $|I|>1$, each point of $D_I\backslash \Sigma$ will be deform to a $|I|-1$ torus under the flow of $V$ and points in $\Sigma$ will not move under the flow of $V$. Let $L_{\infty}\subset D_I \subset X_{\infty}$ be the original fiber of $F$ that move to {Lagrangian } fiber $L\subset X_{\psi}$ under the flow of $V$. Then follow the flow backward we will have a fibration $\pi: L \rightarrow L_{\infty}$. Over $L_{\infty}\backslash \Sigma$, $\pi: L\backslash \Sigma \rightarrow L_{\infty}\backslash \Sigma$ is a $|I|-1$ torus smooth fibration. Over $L_{\infty}\cap \Sigma$, $\pi: L\cap \Sigma \cong L_{\infty}\cap \Sigma$ is an identification.\\\\
It is now easy to see that for $p\in \tilde{\Gamma}^2$, $F^{-1}(p)$ under the flow of $V$ will deform to {Lagrangian } 3-torus with $50$ circles collapsed to $50$ singular points. For $p\in \tilde{\Gamma}^1$, $F^{-1}(p)$ under the flow of $V$ will deform to {Lagrangian } 3-torus with $25$ circles collapsed to $25$ singular points. For $p\in \tilde{\Gamma}^0$, $F^{-1}(p)$ under the flow of $V$ will deform to {Lagrangian } 3-torus with $5$ two-torus collapsed to $5$ singular points. Now we have finished the discussion of all {Lagrangian } fibers of our {Lagrangian } fibration of $X_{\psi}$.\\
\begin{th}
Flow of $V$ will produce a {Lagrangian } fibration $F: X_{\psi} \rightarrow \partial\Delta$. There are 4 types of fibers.\\
(i). For $p\in \partial\Delta\backslash \tilde{\Gamma}$, $F^{-1}(p)$ is a smooth {Lagrangian } 3-torus.\\
(ii). For $p\in \tilde{\Gamma}^2$, $F^{-1}(p)$ is a {Lagrangian } 3-torus with $50$ circles collapsed to $50$ singular points.\\
(iii). For $p\in \tilde{\Gamma}^1$, $F^{-1}(p)$ is a {Lagrangian } 3-torus with $25$ circles collapsed to $25$ singular points.\\
(iv). For $p\in \tilde{\Gamma}^0$, $F^{-1}(p)$ is a {Lagrangian } 3-torus with $5$ two-torus collapsed to $5$ singular points.
\end{th}
\begin{flushright} $\Box$ \end{flushright}
\se{Monodromy of expected {special Lagrangian } fibration}
\subsection{Introduction and assumptions}
In this section we will explore what the SYZ {special Lagrangian } fibration of $X_{\psi}$ should look like if it exists. Recall from Section 2 that $X_{\infty}$ is a union of five ${\bf P^3}$. There is a natural degenerate $T^3$ fibration structure for $X_{\infty}$. Let $\{P_i\}_{i=1}^5$ be five points in ${\bf R^4}$ that are in general position. Consider the natural map $F: \ {\bf P^4}\longrightarrow {\bf R^4}$.\\
\[
F([z])=\sum_{k=1}^5 \frac{|z_k|^2}{\sum_{i=1}^5 |z_i|^2}P_k
\]\\
$\Delta={\rm Image}(F)$ is a 4-simplex. $X_{\infty}$ is naturally fibered over $\partial \Delta$ via this map $F$ with general fiber being $T^3$. This is precisely the $T^3$ special Lagrangian fibration for $X_{\infty}$ as indicated by SYZ construction.\\\\
When $\psi$ is large, we expect that $X_{\psi}$ will also possess a special Largrangian $T^3$ fibration with base identified with $\partial \Delta$ which topologically is an $S^3$. There are still serious analysis and geometric works to be done to totally justify the special Lagrangian fibration. We will discuss in this section that {\bf IF} such special Lagrangian fibration exist on $X_{\psi}$, what should be its expected topological and geometrical structures.\\\\
Our discussion is based on two principle assumptions:\\
(1) {special Lagrangian } fibration on $X_{\psi}$ is a deformation of {special Lagrangian } fibration on $X_{\infty}$. \\
(2) Singular locus of the {special Lagrangian } fibration on $X_{\psi}$ is of codimension 2. \\\\
(1) is very natural, because we expect the {Calabi-Yau } metric on $X_{\psi}$ suitably rescaled (keeping fiber class constant volume) will approach the standard {Calabi-Yau } metric on $X_{\infty}$ when $\psi$ approach infinity. (2) is also very reasonable given the structure of elliptic fibration of complex surface that correspond to dimension $n=2$ situation.\\\\
For $\psi$ large, $X_{\psi}$ will approach $X_{\infty}=\cup_{k=1}^5 D_k$, where $D_k=\{z_k=0\}$. Let us consider the part of the $X_{\psi}$ that is close to $D_5$. (We denote it as $U_5$, for example we may take $U_5$ as inverse image in $X_{\psi}$ of an open set in $D_5$ that stay away from $D_k$ for $k\not=5$ by the following map $\pi_5$.) Then the restriction of the projection\\
\[
\pi_5:\ U_5 \longrightarrow D_5,\ \ \pi_5([z_1,z_2,z_3,z_4,z_5])=[z_1,z_2,z_3,z_4,0]
\]
will identify $U_5$ with an open set in $D_5$, which at the same time will carry over the $T^3$ fibration to $U_5$ from $D_5$. When $\psi$ is large, this fibration should be very close to the special Lagrangian $T^3$ fibration for $U_5$. When we discuss the topology aspect, it will be sufficient to use the induced fibration instead of the special Lagrangian fibration when staying away from intersections of $D_k$'s, where the special Lagrangian fibration of $X_{\infty}$ degenerate. We will use these identifications to compute monodromy of the special Lagrangian fibration in the following.\\
\subsection{One forms and one cycles}
We first fix some notation and introduce some construction. On $D_5$, let $\gamma^1_{52}$ be the circles determined by $\{z_5=0,\ |\frac{z_1}{z_2}| = c_1,\ \frac{z_3}{z_2}=c_2,\ \frac{z_4}{z_2}=c_3\}$. We will also use it to denote circles carried over to $U_5$. In general, we have $\gamma_{ik}^j$ for $\{i,j,k\} \subset \{1,2,3,4,5\}$. Understanding monodromy is equivalent to understanding transformations among $\{\gamma_{ik}^j\}$. It is easy to check that\\
\begin{eqnarray}
\gamma_{ik}^j &=& \gamma_{il}^j,\ \ \ \ \ {\rm for}\ \{i,j,k,l\}\subset \{1,2,3,4,5\}.\label{ba}\\
\gamma_{ik}^j &=& -\sum_{l(\not=i,j)=1}^5\gamma_{ij}^l,\ \ {\rm for}\ \{i,j,k\}\subset \{1,2,3,4,5\}.\nonumber
\end{eqnarray}
On ${\bf P^4}$ we can introduce meromorphic 1-forms\\
\[
\alpha_{ij}= d(\log \frac{z_i}{z_j}), \ \ i \not=j.
\]
They have the simple relations\\
\[
\alpha_{ij}= \alpha_{ik}+ \alpha_{kj}\ {\rm for}\ \{i,j,k\}\subset \{1,2,3,4,5\}.
\]
We also use the same notation for the restriction to $X_{\psi}$. $\alpha_{ij}$ have pole along $D_j\cap X_{\psi}$ on $X_{\psi}$, where\\
\[
D_j\cap X_{\psi}=\left\{z\left| z_j=0,\ \sum_{i(\not=j)=1}^5 z_i^5 = 0\right. \right\}
\]
For this reason we introduce\\
\[
U_j^i = \left\{z\in U_j \left|\ \sum_{k(\not= i,j)=1}^5\left|\frac{z_k}{z_i}\right|^5<1\right. \right\}.
\]
On $U_j^i$, all $\alpha_{kl}$ are regular. On $U_i^j$ we have circles \\
\[
\gamma_{ij}^k, \ {\rm for}\ k\in \{1,2,3,4,5\}\backslash \{i,j\}
\]
and 1-forms\\
\[
\alpha_{kj},\ {\rm for}\ k\in \{1,2,3,4,5\}\backslash \{j\}
\]
It is easy to check that\\
\[
<\gamma_{ij}^k, \alpha_{lj}> = \delta_{kl}\ {\rm for}\ \{i,j,k,l\}\subset \{1,2,3,4,5\}
\]
and\\
\[
<\gamma_{ij}^k, \alpha_{ij}> = -1\ {\rm for}\ \{i,j,k\}\subset \{1,2,3,4,5\}
\]
Use these relations for $U_i^j$ and $U_k^j$, we will get\\
\begin{eqnarray}
\gamma_{ij}^k&=&-\gamma_{kj}^i,\ \ \ \ {\rm for}\ \{i,j,k\}\subset \{1,2,3,4,5\}\label{bb}\\
\gamma_{ij}^l&=&-\gamma_{kj}^i+\gamma_{kj}^l,\ \ {\rm for}\ \{i,j,k,l\}\subset \{1,2,3,4,5\}\nonumber
\end{eqnarray}
From (\ref{ba}) we can see that without confusion, we may denote $\gamma_{ik}^j$ by $\gamma_i^j$. then the relations (\ref{ba}) and (\ref{bb}) can be rewrite together as\\
\begin{eqnarray}
\sum_{j(\not=i)=1}^5\gamma_i^j&=&0,\ \ {\rm for}\ \{i\}\subset \{1,2,3,4,5\}\nonumber\\
\gamma_i^j+\gamma_j^i&=&0,\ \ \ {\rm for}\ \{i,j\}\subset \{1,2,3,4,5\}\label{bc}\\
\gamma_i^j+\gamma_j^k+\gamma_k^i&=&0,\ \ \ {\rm for}\ \{i,j,k\}\subset \{1,2,3,4,5\}\nonumber
\end{eqnarray}
Let\\
\[
U^i = \left\{z\in X_{\psi} \left|\ \sum_{k(\not= i)=1}^5\left|\frac{z_k}{z_i}\right|^5<1\right. \right\}.
\]
Then $\alpha_{ki}$, for $k\in \{1,2,3,4,5\}\backslash \{i\}$ are regular on $U^i$. So there are no monodromy among $U^i_j\subset U^i$ for $j\in \{1,2,3,4,5\}\backslash \{i\}$.\\\\
On the other hand, in $U_j$ the $T^3$ fibration is regular. So there are also no monodromy among $U^i_j\subset U_j$ for $j\in \{1,2,3,4,5\}\backslash \{i\}$. From these discussions we can see that the discriminant locus of the fibration (where the $T^3$ fibration is singular) is topologically a graph $\Gamma$ in $\partial \Delta$. Vertices of $\Gamma$ are $P_{ij}$ (baricenter for 2-simplex $\Delta_{ij}$) for $\{i,j\}\subset \{1,2,3,4,5\}$ and $P_{ijk}$ (baricenter for 3-simplex $\Delta_{ijk}$) for $\{i,j,k\}\subset \{1,2,3,4,5\}$. Legs of $\Gamma$ are $\Gamma_{ij}^k$ which connects $P_{ijk}$ and $P_{ij}$ for $\{i,j,k\}\subset \{1,2,3,4,5\}$.\\\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=4in
\epsffile{p2.ps}}
\end{center}
\caption{}
\end{figure}\\
\subsection{Monodromy around the legs of $\Gamma$}
We would like to compute monodromy $T_{ij}^k$ around $\Gamma_{ij}^k$. To simplify the notation without loss of generality, we will compute $T_{24}^3$. This amounts to compute the transformation:\\
\[
U_5^4\rightarrow U_5^2\rightarrow U_1^2\rightarrow U_1^4\rightarrow U_5^4
\]
Let $S_i^{jk}$ denote the transformation: $U_i^j\rightarrow U_i^k$ and $S^i_{jk}$ denote the transformation: $U^i_j\rightarrow U^i_k$ for $\{i,j,k\}\subset \{1,2,3,4,5\}$. Then\\
\[
T_{24}^3= S_{15}^4 S_1^{24} S_{51}^2 S_5^{42}
\]
For $U_5^4\rightarrow U_5^2$ we have\\
\[
S_5^{42}\left(
\begin{array}{ccc}
\gamma_{54}^1 & \gamma_{54}^2 & \gamma_{54}^3
\end{array}
\right) = \left(
\begin{array}{ccc}
\gamma_{52}^1 & \gamma_{52}^3 & \gamma_{52}^4
\end{array}
\right) \left(
\begin{array}{ccc}
1 & -1 & 0\\
0 & -1 & 1\\
0 & -1 & 0
\end{array}
\right)
\]
For $U_5^2\rightarrow U_1^2$ we have\\
\[
S_{51}^2\left(
\begin{array}{ccc}
\gamma_{52}^1 & \gamma_{52}^3 & \gamma_{52}^4
\end{array}
\right) = \left(
\begin{array}{ccc}
\gamma_{12}^3 & \gamma_{12}^4 & \gamma_{12}^5
\end{array}
\right) \left(
\begin{array}{ccc}
0 & 1 & 0\\
0 & 0 & 1\\
-1& -1& -1
\end{array}
\right)
\]
For $U_1^2\rightarrow U_1^4$ we have\\
\[
S_1^{24}\left(
\begin{array}{ccc}
\gamma_{12}^3 & \gamma_{12}^4 & \gamma_{12}^5
\end{array}
\right) = \left(
\begin{array}{ccc}
\gamma_{14}^2 & \gamma_{14}^3 & \gamma_{14}^5
\end{array}
\right) \left(
\begin{array}{ccc}
0 & -1 & 0\\
1 & -1 & 0\\
0 & -1 & 1
\end{array}
\right)
\]
For $U_1^4\rightarrow U_5^4$ we have\\
\[
S_{15}^2\left(
\begin{array}{ccc}
\gamma_{14}^2 & \gamma_{14}^3 & \gamma_{14}^5
\end{array}
\right) = \left(
\begin{array}{ccc}
\gamma_{54}^1 & \gamma_{54}^2 & \gamma_{54}^3
\end{array}
\right) \left(
\begin{array}{ccc}
-1&-1 &-1\\
1 & 0 & 0\\
0 & 1 & 0
\end{array}
\right)
\]
Therefore\\
\[
T_{24}^3= S_{15}^4 S_1^{24} S_{51}^2 S_5^{42}
\]
\[
=\left(
\begin{array}{ccc}
-1&-1 &-1\\
1 & 0 & 0\\
0 & 1 & 0
\end{array}
\right)
\left(
\begin{array}{ccc}
0 & -1 & 0\\
1 & -1 & 0\\
0 & -1 & 1
\end{array}
\right)
\left(
\begin{array}{ccc}
0 & 1 & 0\\
0 & 0 & 1\\
-1& -1& -1
\end{array}
\right)
\left(
\begin{array}{ccc}
1 & -1 & 0\\
0 & -1 & 1\\
0 & -1 & 0
\end{array}
\right)
\]
\[
=\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]
and\\
\[
T_{24}^3 \left(
\begin{array}{ccc}
\gamma_{54}^1 & \gamma_{54}^2 & \gamma_{54}^3
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{54}^1 & \gamma_{54}^2 & \gamma_{54}^3
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]
In general for $\{i,j,k,l,m\}=\{1,2,3,4,5\}$ representing the same orientation with the written order\\
\[
T_{jl}^k \left(
\begin{array}{ccc}
\gamma_{m}^i & \gamma_{m}^j & \gamma_{m}^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{m}^i & \gamma_{m}^j & \gamma_{m}^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right),
\]\\
where the monodromy is computed along the path\\
\[
U_m^l\rightarrow U_m^j\rightarrow U_i^j\rightarrow U_i^l\rightarrow U_m^l.
\]\\
If orientation is different, then $-5$ should be replaced by $5$.\\
\subsection{Monodromy around baricenter of 2-cell}
Now we compute the monodromy around a baricenter of a 2-cell.\\\\
Let $S_m^{kl}$ denote the transformation: $U_m^k\rightarrow U_m^l$\\
\[
S_m^{kl}\left(
\begin{array}{ccc}
\gamma_m^i & \gamma_m^j & \gamma_m^l
\end{array}
\right) = \left(
\begin{array}{ccc}
\gamma_m^i & \gamma_m^j & \gamma_m^k
\end{array}
\right) \left(
\begin{array}{ccc}
1 & 0 & -1\\
0 & 1 & -1\\
0 & 0 & -1
\end{array}
\right)
\]\\
and $S^k_{ml}$ denote the transformation: $U^k_m\rightarrow U^k_l$\\
\[
S_{ml}^k\left(
\begin{array}{ccc}
\gamma_m^i & \gamma_m^j & \gamma_m^l
\end{array}
\right) = \left(
\begin{array}{ccc}
\gamma_l^i & \gamma_l^j & \gamma_l^m
\end{array}
\right) \left(
\begin{array}{ccc}
1 & 0 & 0\\
0 & 1 & 0\\
-1& -1& -1
\end{array}
\right)
\]\\
For $(i,j,k,l,m)=(1,4,2,3,5)$\\
\[
T_{43}^2 \left(
\begin{array}{ccc}
\gamma_5^1 & \gamma_5^4 & \gamma_5^2
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_5^1 & \gamma_5^4 & \gamma_5^2
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]\\
For $(i,j,k,l,m)=(1,3,4,2,5)$\\
\[
T_{32}^4 \left(
\begin{array}{ccc}
\gamma_5^1 & \gamma_5^3 & \gamma_5^4
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_5^1 & \gamma_5^3 & \gamma_5^4
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]\\
Write under the same basis\\
\[
T_{43}^2 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=
T_{43}^2 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^4
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & -1\\
0 & 1 & -1\\
0 & 0 & -1
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^4
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & -5\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & -1\\
0 & 1 & -1\\
0 & 0 & -1
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & -1\\
0 & 1 & -1\\
0 & 0 & -1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & -5\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & -1\\
0 & 1 & -1\\
0 & 0 & -1
\end{array}
\right)
\]\\
which gives\\
\[
T_{43}^2 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & 5\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]\\
\[
T_{32}^4 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=
T_{32}^4 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^3 & \gamma_{5}^4
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -1 & 0\\
0 & -1 & 1\\
0 & -1 & 0
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^3 & \gamma_{5}^4
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -1 & 0\\
0 & -1 & 1\\
0 & -1 & 0
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & -1\\
0 & 0 & -1\\
0 & 1 & -1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -1 & 0\\
0 & -1 & 1\\
0 & -1 & 0
\end{array}
\right)
\]\\
which gives\\
\[
T_{32}^4 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 5 & -5\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]\\
From these expressions, it is clear that $T_{32}^4$, $T_{24}^3$ and $T_{43}^2$ commute with each other and\\
\[
T_{32}^4 T_{24}^3 T_{43}^2= Id.
\]\\
They determine a natural filtration:\\
\[
{\cal W}^0 \subset {\cal W}^1 = H_1(T^3, {\bf Z})
\]\\
with ${\cal W}^0$ generated by the vanishing circle $\gamma_{5}^1$, which is the common vanishing cycle of $T_{32}^4$, $T_{24}^3$ and $T_{43}^2$. Recall that the $T^3$ fibration of $X_{\infty}$ over $\partial \Delta$ degenerate to be a $T^2$ fibration over $\Delta _{15}$. An interesting fact is that this $T^2$ is exactly the quotient of $T^3$ by $\gamma_{5}^1$.\\
\subsection{Monodromy around baricenter of 1-cell}
We can similarly compute the monodromies around a midpoint of a 1-edge, for instance, $P_{24}$. We have for $(i,j,k,l,m)=(5,2,1,4,3)$\\
\[
T_{24}^1 \left(
\begin{array}{ccc}
\gamma_{3}^5 & \gamma_{3}^2 & \gamma_{3}^1
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{3}^5 & \gamma_{3}^2 & \gamma_{3}^1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]\\
and for $(i,j,k,l,m)=(3,2,5,4,1)$\\
\[
T_{24}^5 \left(
\begin{array}{ccc}
\gamma_{1}^3 & \gamma_{1}^2 & \gamma_{1}^5
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{1}^3 & \gamma_{1}^2 & \gamma_{1}^5
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)
\]\\
write under the same basis $(\gamma_{5}^1, \gamma_{5}^2, \gamma_{5}^3)$.\\
\[
T_{24}^1 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=
T_{24}^1 \left(
\begin{array}{ccc}
\gamma_{3}^5 & \gamma_{3}^2 & \gamma_{3}^1
\end{array}
\right)\left(
\begin{array}{ccc}
-1 & -1 & -1\\
0 & 1 & 0\\
1 & 0 & 0
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{3}^5 & \gamma_{3}^2 & \gamma_{3}^1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
-1&-1 & -1\\
0 & 1 & 0\\
1 & 0 & 0
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
0 & 0 & 1\\
0 & 1 & 0\\
-1& -1 &-1
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
-1&-1 & -1\\
0 & 1 & 0\\
1 & 0 & 0
\end{array}
\right)
\]\\
which gives\\
\[
T_{24}^1 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & 0\\
0 & 1 & 0\\
0 & 5 & 1
\end{array}
\right)
\]\\
\[
T_{24}^5 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=
T_{24}^5 \left(
\begin{array}{ccc}
\gamma_{1}^3 & \gamma_{14}^2 & \gamma_{1}^5
\end{array}
\right)\left(
\begin{array}{ccc}
0 & 0 & 1\\
0 & 1 & 0\\
-1 & -1 & -1
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{1}^3 & \gamma_{1}^2 & \gamma_{1}^5
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
0 & 0 & 1\\
0 & 1 & 0\\
-1 & -1 & -1
\end{array}
\right)
\]\\
\[
=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
-1& -1 &-1\\
0& 1 & 0\\
1& 0 & 0
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right)\left(
\begin{array}{ccc}
0 & 0 & 1\\
0 & 1 & 0\\
-1 & -1 & -1
\end{array}
\right)
\]\\
which gives\\
\[
T_{24}^5 \left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{5}^1 & \gamma_{5}^2 & \gamma_{5}^3
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 5 & 0\\
0 & 1 & 0\\
0 & -5 & 1
\end{array}
\right)
\]\\
From these expressions, it is clear that $T_{24}^1$, $T_{24}^3$ and $T_{24}^5$ commute with each other and\\
\[
T_{24}^1 T_{24}^3 T_{24}^5= Id.
\]\\
They determine a natural filtration:\\
\[
{\cal W}^0 \subset {\cal W}^1 = H_1(T^3, {\bf Z})
\]\\
with ${\cal W}^0$ generated by the vanishing circle $\gamma_{5}^1$ and $\gamma_{5}^3$. Recall that the $T^3$ fibration of $X_{\infty}$ over $\partial \Delta$ degenerate to be a $T^1$ fibration over $\Delta _{135}$. An interesting fact is that this $T^1$ is exactly the quotient of $T^3$ by $\gamma_{5}^1\times \gamma_{5}^3$.\\\\
\se{Geometry of the singular fibers}
With monodromy information in mind, we would like to discuss possible structure of singular special Lagrangian fibers. We will start by describing a possible model for the structure of {special Lagrangian } fibration, especially the singular fibers, that is of conjectural nature. Then we will use our construction of {Lagrangian } fibration to give some partial justification.\\\\
Suppose that we have a {special Lagrangian } fibration $F: X_{\psi} \rightarrow \partial \Delta$. Then the monodromy information will suggest that it is likely that $F$ is smooth fibration over $\partial \Delta\backslash \Gamma$, where the singular locus $\Gamma$ is topologically a graph in $\partial \Delta$. Vertices of $\Gamma$ are $P_{ij}$ (baricenter for 2-simplex $\Delta_{ij}$) for $\{i,j\}\subset \{1,2,3,4,5\}$ and $P_{ijk}$ (baricenter for 3-simplex $\Delta_{ijk}$) for $\{i,j,k\}\subset \{1,2,3,4,5\}$. Legs of $\Gamma$ are $\Gamma_{ij}^k$ which connects $P_{ijk}$ and $P_{ij}$ for $\{i,j,k\}\subset \{1,2,3,4,5\}$.\\\\
In general for $\{i,j,k,l,m\}=\{1,2,3,4,5\}$ representing the same orientation with the written order, the monodromy around $\Gamma_{ij}^k$ computed along the path\\
\[
U_m^j\rightarrow U_m^i\rightarrow U_l^i\rightarrow U_l^j\rightarrow U_m^j.
\]\\
can be written as\\
\[
T_{ij}^k \left(
\begin{array}{ccc}
\gamma_{m}^l & \gamma_{m}^i & \gamma_{m}^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_{m}^l & \gamma_{m}^i & \gamma_{m}^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right).
\]\\
This suggests that singular fibers on $\Gamma_{ij}^k$ should be of $I_5 \times S^1$ type, where $I_5$ is a Kodaira type elliptic singular fiber as degeneration of $\gamma_{m}^l \times \gamma_{m}^i$ with vanishing cycle $\gamma_{m}^l$, and the $S^1$ corresponds to $\gamma_{m}^k$. This is a general type of singular fibers. Let's call it type $I_5$ singular fiber. Since Type $I_5$ singular fibers have $S^1$ factor, they give no contribution to Euler number.\\\\
On the other hand, around $P_{ijk}$, monodromies are\\
\[
T_{ij}^k \left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right).
\]
\[
T_{jk}^i \left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & 5\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right).
\]
\[
T_{ki}^j \left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 5 &-5\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right).
\]
They determine the natural filtration:\\
\[
{\cal W}^0 \subset {\cal W}^1 = H_1(T^3, {\bf Z})
\]
with ${\cal W}^0$ generated by the common vanishing cycle $\gamma_m^l$. By the symmetric property, the singular fiber over $P_{ijk}$ can be described as follows. Let $\Gamma_{ijk}$ be the graph in $T^2\cong T^3/\gamma_{m}^l$ as discribed by the following picture, where $\{lm\}=\overline{\{ijk\}}$. Then the singular fiber over $P_{ijk}$ can be identified topologically as $T^3$ collapsing circles $(\gamma_{m}^l)$ over $\Gamma_{ijk}$ to points. We call this kind of fibers type $II_{5\times 5}$. It is easy to see that the Euler number of a type $II_{5\times 5}$ fiber is $-25$.\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=3in
\epsffile{six2.eps}}
\end{center}
\caption{Type $II_{5\times 5}$ fiber}
\end{figure}\\
Finally we are left to determine the singular fiber over $P_{ij}$. Around $P_{ij}$ the monodromies are\\
\[
T_{ij}^k \left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & -5 & 0\\
0 & 1 & 0\\
0 & 0 & 1
\end{array}
\right).
\]
\[
T_{ij}^l \left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 0 & 0\\
0 & 1 & 0\\
0 & 5 & 1
\end{array}
\right).
\]
\[
T_{ij}^l \left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)=\left(
\begin{array}{ccc}
\gamma_m^l & \gamma_m^i & \gamma_m^k
\end{array}
\right)\left(
\begin{array}{ccc}
1 & 5 & 0\\
0 & 1 & 0\\
0 & -5 & 1
\end{array}
\right).
\]
They determine the natural filtration:\\
\[
{\cal W}^0 \subset {\cal W}^1 = H_1(T^3, {\bf Z})
\]
with ${\cal W}^0$ generated by the vanishing circles $\gamma_m^l$ and $\gamma_m^k$. The singular fiber over $P_{ij}$ is sort of generalization of $I_5$ singularity to $3$-dimension. Topologically equivalent to $T^3$ collapsing five of sub-$T^2$ equivalent to $\gamma_m^l\times \gamma_m^k$. It looks like five $3$-dimensional pseudomanifolds linked into a ``necklace'' via their singular points. Each $3$-dimensional pseudomanifold is a $T^2\times [0,1]$ collapsing the two boundaries to two singular points. An illustration of this singular fiber is in the following picture. We will call these type $III_5$ singular fibers. A type $III_5$ singular fiber has Euler number $5$.\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=4in
\epsffile{p3.eps}}
\end{center}
\caption{Type $III_5$ fiber}
\end{figure}\\
All together we have $10$ type $II_{5\times 5}$ and $10$ type $III_5$ singular fibers, which give total Euler number\\
\[
\chi(X)=10\times (-25)+10\times 5 = -200.
\]
This is exactly the Euler number of a smooth quintic {Calabi-Yau } threefold.\\\\
It is instructive to understand the singular point set of the fibration map $f$, namely the points in $X_{\psi}$ where tangent maps of $f$ are not surjective. The singular point set of a type $I_5$ fiber is five $S^1$'s. The singular point set of a type $II_{5\times 5}$ fiber is the graph in figure 2. The singular point set of a type $III_5$ fiber is five points. Together they form a Riemann surface. The Riemann surface has $10$ irreducable components $\{\Sigma_{ijk}\}$. $\Sigma_{ijk}$ is fibered over $\Gamma_{ij}^k\cup \Gamma_{ki}^j\cup \Gamma_{jk}^i$ and centered over $P_{ijk}$. Singular fibers are three of $5$-point sets and $\Gamma_{ijk}$. Total Euler number is \\
\[
\chi(\Sigma_{ijk})=3\times 5 - 25 = -10 = 2-2g(\Sigma_{ijk}).
\]
So $g(\Sigma_{ijk})= 6$.\\\\
Actually there is a very explicit description of the local structure of the map $f$ arround a singular point in a type $III_5$ fiber. It was given in the celebrated paper of R. Harvey and H. B. Lawson (\cite{HL}). The construction is as follows.\\
\begin{th}
Let $M_c=f^{-1}(c)$, where $f:\ {\bf C^3}\longrightarrow {\bf R^3}$\\
\[
f_i(z)= |z_1|^2 - |z_i|^2,\ \ i=2, 3;\ \ \ f_1= {\rm Im}(z_1 z_2 z_3).
\]
Then $M_c$ (with the correct orientation) is a special Lagrangian submanifold of ${\bf C^3}$.
\end{th}
\begin{flushright} $\Box$ \end{flushright}
It is pretty obvious why this characterize the singular points in type $III_5$ singular fiber. Since each $M_c$ is invariant under the action of the group\\
\[
T^2=\left(
\begin{array}{ccc}
e^{i\theta_1} & 0 & 0\\
0 & e^{i\theta_2} & 0\\
0 & 0 & e^{i\theta_3}
\end{array}
\right)\ \ \theta_1+\theta_2+\theta_3=0.
\]
An easy computation will show that the singular point set around origin is exactly the union of the three coordinate axes. This is where three of the irreducible components of the singular point set meet. The singular locus in ${\bf R^3}$ is\\
\[
\{c\in{\bf R^3}| c_2\leq 0, c_3=c_1=0\}\cup \{c\in{\bf R^3}| c_3\leq 0, c_2=c_1=0\}
\]
\[
\cup \{c\in{\bf R^3}| c_2+c_3\geq 0, c_2-c_3=c_1=0\}.
\]
Fiber over origin is of type $III_5$ singularity and fibers over other points in the singular locus are Type $I_5$. An important observation is that $\Sigma_{ijk}$ is exactly equivalent to the genus 6 curve\\
\[
\{[z]\in {\bf CP^4}| z_i^5 + z_j^5 + z_k^5 =0, z_l=0\ \ {\rm for}\ l\in\{1,2,3,4,5\}\backslash\{i,j,k\}\}
\]
Associate to the special lagrangian torus fiberation $f$ there is the Leray spectral sequence, which abuts to $H^\cdot(X_{\psi}, {\bf Q})$ and in which\\
\[
E_2^{p,q} = {\bf H}^p(\partial \Delta; R^qf_*{\bf Q})
\]
This spectral sequence degenerates at $E_2$ term. We can use it to compute cohomology of $X_{\psi}$.\\\\
This spectral sequence was discussed and computed in \cite{GW} for expected SYZ {special Lagrangian } fiberation of generic {Calabi-Yau } manifolds. Our situation is far from generic. Fermat type quintics have a lot of symmetries and actually represent singular (orbifold) points in the moduli space of {Calabi-Yau }. It is interesting to compute the spectral sequence of the fiberation and compare with the generic situation.\\\\
Since $R^0f_*{\bf Q} ={\bf Q}$, $\partial \Delta \cong S^3$. It is easy to see that the bottom row of the spectral sequence is\\
\[
E_2^{3,q}\cong H^q(S^3,{\bf Q}):\ \ \ {\bf Q},\ 0,\ 0,\ {\bf Q}.
\]
In general, we have the following exact sequence\\
\[
0\longrightarrow K_q \longrightarrow R^qf_*{\bf Q} \longrightarrow i_* E_q \longrightarrow 0
\]
with the second non-trivial map being the attachment map with $E_q=i^* R^qf_*{\bf Q}$ for $i:\ \partial\Delta\backslash \Gamma \longrightarrow \partial\Delta$. Space $B=\partial\Delta$ has a natural filtration\\
\[
B=B_3\supset B_2\supset B_1\supset B_0\supset B_0 = \phi
\]
where $B_2=B_1=\Gamma$, $B_0=\{P_{ij}|i,j\in \{1,2,3,4,5\}\}\cup \{P_{ijk}|i,j,k\in \{1,2,3,4,5\}\}$. Each stratum $S_i=B_i\backslash B_{i-1}$ is a $i$-manifold. We denote $V_i=B\backslash B_i$. $E_q$ is a local system on $V_1=\partial\Delta\backslash \Gamma$ and $K_q$ is supported on $B_1=\Gamma$. Each $R^qf_*{\bf Q}$ is a locally constant sheaf. Its germ over $P\in B$ is\\
\[
(R^qf_*{\bf Q})_P\cong H^q(f^{-1}(P), {\bf Q}) \cong (H_q(f^{-1}(P), {\bf Q}))^\vee
\]
$E_3$ is obviously a constant sheaf, so $i_*E_3\cong {\bf Q}$. According to our geometric description of the singular fibers, $R^3f_*{\bf Q}$ is of dimension $1$ over $S_3$, dimension $5$ over $S_1$, dimension $5$ over $P_{ij}$ and dimension $25$ over $P_{ijk}$. Therefore $K_3$ is supported in $B_1=\Gamma$ and is of dimension $4$ over $S_1$, dimension $4$ over $P_{ij}$ and dimension $24$ over $P_{ijk}$. Clearly only possible non-trivial cohomologies for $K_3$ are $H^0(K_3,{\bf Q})$ and $H^1(K_3,{\bf Q})$.\\\\
\begin{prop}
\[
H^1(K_3,{\bf Q})=0
\]
\end{prop}
{\bf Proof:} We will compute the {\v C}ech cohomology. There is a natural map\\
\[
\pi_{ijk}:\ (K_3)_{P_{ijk}} \longrightarrow (K_3)_{\Gamma^k_{ij}}\oplus (K_3)_{\Gamma^i_{jk}}\oplus (K_3)_{\Gamma^j_{ki}}
\]
We will show that $\pi_{ijk}$ is surjective, which easily implies that $H^1(K_3,{\bf Q})=0$. Observe that $\pi_{ijk}$ is induced from\\
\[
{\tilde \pi}_{ijk}: (R^3f_*{\bf Q})_{P_{ijk}} \longrightarrow (R^3f_*{\bf Q})_{\Gamma^k_{ij}}\oplus (R^3f_*{\bf Q})_{\Gamma^i_{jk}}\oplus (K_3)_{\Gamma^j_{ki}}
\]
Choose generators $\{x_{ij}|i,j\in \{1,2,3,4,5\}\}$ of $(R^3f_*{\bf Q})_{P_{ijk}}$, $\{u_i\}_{i=1}^5$ of $(R^3f_*{\bf Q})_{\Gamma^k_{ij}}$, $\{v_i\}_{i=1}^5$ of $(R^3f_*{\bf Q})_{\Gamma^j_{ki}}$ and $\{w_i\}_{i=1}^5$ of $(R^3f_*{\bf Q})_{\Gamma^k_{ij}}$ (representing geometric cycles). If we choose correctly, we should have\\
\[
{\tilde \pi}_{ijk}(x_{lm})=(u_l, v_m, w_n),\ \ n \equiv -l-m\ (mod\ 5).
\]
By simple linear algebra, we can see that\\
\[
\left( \sum^5_{i=1}a_l u_l,\ \sum^5_{i=1}b_l v_l,\ \sum^5_{i=1}c_l w_l\right) \in {\rm Im}({\tilde \pi}_{ijk})
\]
if and only if
\[
\sum^5_{i=1}a_l =\sum^5_{i=1}b_l =\sum^5_{i=1}c_l;
\]
and
\[
\left( \sum^5_{i=1}a_l u_l,\ \sum^5_{i=1}b_l v_l,\ \sum^5_{i=1}c_l w_l\right) \in {\rm Im}(\pi_{ijk})
\]
if and only if
\[
\sum^5_{i=1}a_l =\sum^5_{i=1}b_l =\sum^5_{i=1}c_l =0.
\]
Therefore $\pi_{ijk}$ is surjective.
\begin{flushright} $\Box$ \end{flushright}
\begin{prop}
\[
h^0(K_3)=160.
\]
\end{prop}
{\bf Proof:} We will compute Euler number $e(K_3)=h^0(K_3)-h^1(K_3)$. We compute in the chain level. Let $c^i=\dim(C^i)$ be the dimension of the space of dimension $i$ cochain. It is easy to see\\
\[
c^0=10\times 24 + 10 \times4 =280,\ \ c^1=30\times4=120,\ \ e(K_3)=c^0-c^1=160.
\]
Then $h^1(K_3)=0$ from last proposition gives us that $h^0(K_3)=160.$
\begin{flushright} $\Box$ \end{flushright}
Now use the long exact sequence associated to\\
\[
0\longrightarrow K_3 \longrightarrow R^3f_*{\bf Q} \longrightarrow i_* E_3 \longrightarrow 0
\]
we get
\begin{prop}
\[
h^0(R^3f_*{\bf Q})=161,\ h^1(R^3f_*{\bf Q})=h^2(R^3f_*{\bf Q})=0,\ h^3(R^3f_*{\bf Q})=1.
\]
\end{prop}
\begin{flushright} $\Box$ \end{flushright}
To compute the cohomology for $R^1f_*{\bf Q}$, notice that $R^1f_*{\bf Q}=i_*E_1$ for local system $E_1$. It is easy to see that the monodromy of the local system $E_1$ generate $SL_3({\bf Z})$. So $h^0(R^1f_*{\bf Q})=0$. From this we will have\\
\begin{prop}
\[
h^3(R^1f_*{\bf Q})=h^3(R^2f_*{\bf Q})=0.
\]
\end{prop}
{\bf Proof:} By Poincare duality for general fibers, $E_2=E_1^\vee$. So\\
\[
R^2f_*{\bf Q}\cong {\cal D}i_*E_1 \ \ \ {\rm in}\ V_2
\]
where ${\cal D}E$ is the dual of $E$ in the sense of Verdier. Then by Verdier duality\\
\[
H^3(R^2f_*{\bf Q})\cong H^3({\cal D}i_*E_1)\cong H^0(i_*E_1)^\vee=0.
\]
The space $B$ have a natural symmetry $s$ which respects the filtration. It is a piecewise linear map satisfying\\
\begin{equation}
\label{ca}
s^{-1}=s:\ B \longrightarrow B;\ s(P_i)=P_{\overline{i}},\ s(P_{ij})=P_{\overline{ij}}
\end{equation}
where $\overline{S}$ indicate compliment of $S\in \{1,2,3,4,5\}$. It is easy to check by looking at the monodromy that $s^*E_1\cong E_2$. (Later we will see that this fact is related to the mirror symmetry construction of Fermat type mirror construction.) Therefore\\
\[
i_*E_1\cong s^*R^2f_*{\bf Q}\ \ \ {\rm in}\ V_2
\]
and
\[
H^3(R^1f_*{\bf Q})=H^3(i_*E_1)\cong H^3(R^2f_*{\bf Q})=0.
\]
\begin{flushright} $\Box$ \end{flushright}
To compute $H^1(R^1f_*{\bf Q})$, we need to use intersection cohomology. Since $R^1f_*{\bf Q}=i_*E_1$, We have $H^1(R^1f_*{\bf Q})=I_0 H^1(B,E_1)$. Where $I_p H^i(B,E)$ denote the $i$-th intersection cohomology of $B$ with coefficient in the local system $E$ and perversity $p$.\\
\begin{prop}
\label{bd}
\[
h^1(R^1f_*{\bf Q})=1.
\]
\end{prop}
{\bf Proof:} By Poincare duality and $E_2=E_1^\vee$, we have\\
\[
H^1(R^1f_*{\bf Q})=I_0 H^1(B,E_1)=I_t H^2(B,E_2)^\vee
\]
where $t$ denote the top perversity. Intersection homology is related to intersection cohomology as\\
\[
I_t H^2(B,E_2)=I_t H_1(B,E_2).
\]
So we only need to compute $I_t H_1(B,E_2)$, where for $P\in V_1$ we have $(E_2)_P\cong H_1(f^{-1}(P),{\bf Q})$. Consider the first barycentric subdivision of the standard triangulation of $B$. The 1-simpleces allowed to compute $I_t H_1(B,E_2)$ are the 1-simpleces $\Gamma^i_{jkl}$ connecting $P_i$ and $P_{ijkl}$.\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=5.5in
\epsffile{p4a.eps}}
\end{center}
\caption{}
\end{figure}\\
We can introduce the following 1-chain\\
\[
L=\gamma_j^i \cdot \Gamma^i_{\overline{ij}}
\]
By (\ref{bc}), it is easy to see that $\partial L=0$. Actually this $L$ generate $I_t H_1(B,E_2)$, namely\\
\[
h^1(R^1f_*{\bf Q}) = \dim (I_t H_1(B,E_2))=1
\]
\begin{flushright} $\Box$ \end{flushright}
\begin{prop}
\label{be}
\[
h^2(R^1f_*{\bf Q})=1.
\]
\end{prop}
{\bf Proof:} We will show this by proving that $\chi (R^1f_*{\bf Q})=0$, which clearly implies this proposition by results in proposition \ref{bd} and \ref{be}. The Euler number can be computed in two ways, either by straightforward {\v C}ech cohomology computation or by the identification\\
\[
H^i(R^1f_*{\bf Q}) \cong I_0 H^i(B,E_1)\cong I_0 H_{3-i}(B,E_1).
\]
We will use the latter approach, which is more elegant. To get the number right, the key point is to choose the right triangulation, which respects the filtration of $B$. The previous trangulation as indicated in figure 4 is not good enough. The problem is that the 0-simplices $P_{ij},P_{ijk}\in B_0$ will span a 1-simplex $\Gamma^k_{ij}$ which is in $B_1$ but not in $B_0$. The right triangulation which respects the filtration of $B$ can be achieved by taking the baricentric subdivision of $\Gamma$ and then naturally extend to a subdivision of the previous triangulation of $B$. Practically, each old 3-simplex is divided into two new 3-simleces by a new 2-simplex as indicated in the following picture.\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=5.5in
\epsffile{p5b.eps}}
\end{center}
\caption{}
\end{figure}\\
It is easy to see that\\
\[
I^0C_0=\bigoplus_{i=1}^5(E_1)_{P_{i}}\cdot P_{i} \oplus (E_1)_{P_{\overline{i}}}\cdot P_{\overline{i}}
\]
\[
I^0C_1=\bigoplus_{\tiny{\begin{array}{c} i,j=1\\ (i\not= j)\end{array}}}^5(E_1)_{<P_{i},P_{\overline{j}}>}\cdot <P_{i},P_{\overline{j}}>,
\]
and $c_0=10\times 3=30$, $c_1=20\times 3=60$.\\\\
For the $I^0C_2$, it is easy to see that only the newly added 2-simplices $<P_{i}, P_{\{ij\}\{ijk\}}, P_{ijkl}>$ will be relavent. Since the boundary of the intersection 2-chain should not intersect $\Gamma$. An intersection 2-chain containing $<P_{i}, P_{\{ij\}\{ijk\}}, P_{ijkl}>$ should also contain all other newly added 2-simplices that contain $P_{\{ij\}\{ijk\}}$, namely should contain\\
\[
\Delta^k_{ij}=\bigcup\{{\rm 2-simplex\ with}\ P_{\{ij\}\{ijk\}}\ {\rm as\ one\ vertex}\}.
\]
Therefore\\
\[
I^0C_2=\bigoplus_{\Gamma_{ij}^k}(E_1)^{T^k_{ij}}_{\Delta^k_{ij}}\cdot \Delta^k_{ij}
\]
where the upper index ${T^k_{ij}}$ indicates the ${T^k_{ij}}$ invariant piece of $E_1$. So $c_2=30\times2=60$.\\\\
Since an intersection 2-chain is not supposed to contain $P_{ij}$ or $P_{ijk}$. The boundary of an intersection 3-chain also should not contain $P_{ij}$ or $P_{ijk}$. Hence an intersection 3-chain containing $<P_{i}, P_{ij}, P_{\{ij\}\{ijk\}}, P_{ijkl}>$ has to contain\\
\[
\Delta_{ij}=\bigcup\{{\rm 3-simplex\ with}\ P_{ij}\ {\rm as\ one\ vertex}\}.
\]
An intersection 3-chain containing $<P_{i}, P_{ijk}, P_{\{ij\}\{ijk\}}, P_{ijkl}>$ has to contain\\
\[
\Delta_{ij}=\bigcup\{{\rm 3-simplex\ with}\ P_{ijk}\ {\rm as\ one\ vertex}\}.
\]
And\\
\[
I^0C_3=\left(\bigoplus_{ij} (E_1)^{G_{ij}}_{\Delta_{ij}}\cdot \Delta_{ij}\right) \bigoplus \left(\bigoplus_{ijk} (E_1)^{G_{ijk}}_{\Delta_{ijk}}\cdot \Delta_{ijk}\right),
\]
where $G_{ij}$ is the monodromy group around $P_{ij}$ and $G_{ijk}$ is the monodromy group around $P_{ijk}$. And $c_3=10\times 1+10\times2=30$. Therefore\\
\[
\chi(R^1f_*{\bf Q})=-\chi(I^0C.)=-\sum_{i=0}^3 (-1)^i c_i = -(30-60+60-30)=0.
\]
\begin{flushright} $\Box$ \end{flushright}
Now the only thing we do not yet know is $h^1(R^2f_*{\bf Q})$.\\
\begin{prop}
\[
h^1(R^2f_*{\bf Q})=41.
\]
\end{prop}
{\bf Proof:} Clearly we only need to compute $h^1(K_2)$. Since the germ of $K_2$ at $P_{ij}$ only have zero section. It is clear that $h^0(K_2)=0$. It is now reduced to compute $\chi(K_2)$. We will simply count the {\v C}ech cochains.\\
\[
c^1=(5-1)\times 30=120,\ \ c^0=(10-2)\times 10 = 80.
\]
\[
\chi(K_2)=c^0-c^1 = 80-120=-40.
\]
So $h^1(K_2)=40$ and $h^1(R^2f_*{\bf Q})=41$.
\begin{flushright} $\Box$ \end{flushright}
Now we have the whole $E_2$ term of the spectral sequence.\\
\begin{th}
$E_2$ term of the Leray spectral sequence of fibration $F$ is\\
\[
\begin{array}{cccc}
161 & 0 & 0 & 1\\
0 & 41 & 1 & 0\\
0 & 1 & 1 &0\\
1 & 0 & 0 & 1
\end{array}
\]
\end{th}
\begin{flushright} $\Box$ \end{flushright}
One can see that the $E_2$ term of our situation is quite different from the $E_2$ term of generic {Calabi-Yau } situation as indicated in \cite{GW}.\\\\
Note: Our discussion so far is of conjectural nature. Since we can not construct {special Lagrangian } fibration, we can only guess its possible structure based on our monodromy discussion.\\
\se{Comparison}
The {Lagrangian } fibration we constructed in Section 3 is different from the expected {special Lagrangian } fibration structure described in the previous two sections. Yet they are very closely related. The singular locus $\Gamma$ of the expected {special Lagrangian } fibration is of codimension 2 (an one dimensional graph) and each singular fiber has singularity of codimension 2. The singular locus $\tilde{\Gamma}$ of the {Lagrangian } fibration we constructed is of codimension 1 and most singular fibers have singularity of codimension 3. On the other hand, $\tilde{\Gamma}$ is just a fattened version of $\Gamma$. $S^3\backslash\tilde{\Gamma}$ and $S^3\backslash\Gamma$ have the identical fundamental group. We can compare monodromies of the two fibrations.\\\\
Since our {Lagrangian } fibration on $X_{\psi}$ is a deformation of the standard {special Lagrangian } fibration on $X_{\infty}$, it satisfies the principle assumption (1) that we based to compute monodromy. For this reason, it is not hard to see that above computation of monodromy naturally apply to our {Lagrangian } fibration. Therefore\\
\begin{th}
Our {Lagrangian } fibration have the same monodromy as the monodromy computed for the expected {special Lagrangian } fibration.
\end{th}
\begin{flushright} $\Box$ \end{flushright}
The singularities of the two fibrations are also closely related. For one thing, the singular point set of our construction in $X_{\psi}$ is $\Sigma$ (the union of 10 genus 6 curves), and the singular point set of the expected {special Lagrangian } fiberation is naturally equivalent to $\Sigma$. One way to see this is to notice that we can easily construct a homotopy contraction $p: \tilde{\Gamma} \rightarrow \Gamma$ that is the homotopy inverse of the natural injection $i: \Gamma \rightarrow \tilde{\Gamma}$ that satisfy $p\cdot i= id$ (for instance map point in $\tilde{\Gamma}$ to the nearest point in $\Gamma$). Let $q$ denote the composition of projection of $\Sigma$ to $\tilde{\Gamma}$ and $p$. Then $q$ map $\Sigma$ to $\Gamma$. It is easy to observe that the inverse image of a point $P$ in $\Gamma$ with respect to $q$ is exactly corresponding to singular set of the expected singular {special Lagrangian } fiber over $P$ as described in the previous section. With this fact in mind, we intend to modify our {Lagrangian } torus fibration into the expected topological shape.\\\\
If one is willing to sacrifies {Lagrangian } property, it is not hard to deform the fibration we constructed into a non-{Lagrangian } smooth fiberation with the same topological shape as we expected of {special Lagrangian } fibration. One way to do this is to notice that there is a natural horizontal foliation in ${\bf P^2}$ where one of the component of $\Sigma$ is located. Then conceivably, one can smoothly deform part of 2-torus above $\tilde{\Gamma}$ that intersect $\Sigma$ along horizontal direction away from $\Gamma$ in direction indicated by map $p$. So that eventually only those 2-torus above $\Gamma$ will intersect $\Sigma$ at one dimensional graphs that is the inverse image under $q$ of the corresponding point. Then we run the flow of $V$, we will get a non-{Lagrangian } torus fibration with topological structure that is the same as expected {special Lagrangian } fibration. This will give an alternative simple proof of the result in \cite{Z} that gave a smooth torus non-{Lagrangian } fibration. Since our major concern is to construct smooth {Lagrangian } torus fiberation, we will not get into great details in this direction.\\\\
However to get a {Lagrangian } torus fibration of the right shape is much trickier. We will address this problem in the sequel to this paper(\cite{lag2}).\\\\
Another important observation is that for an $n$-torus {Lagrangian } fibration such that the fibration map is $C^{\infty}$ (actually $C^2$ is enough), there is a natural action of ${\bf R^n}$ on each fiber (smooth or singular). Look at our construction, one can see that there are no action of ${\bf R^3}$ on most of our singular fibers, which seems to be a contradiction. It turns out that fibration map of our construction is not $C^{\infty}$---it is typically only piecewise smooth and globally merely $C^{0,1}$ (Lipschitz). This reveals a major difference between a non-{Lagrangian } fibration and a {Lagrangian } fibration. If one does not care keeping the {Lagrangian } property, the map usually can be smoothed by a small perturbation. Above ovservation shows that in general a {Lagrangian } fiberation can not be deformed to a smooth {Lagrangian } fiberation by small perturbation. In (\cite{lag2}) we will explore this issue further and try to get a smooth {Lagrangian } fibration. We will also discuss general construction of {Lagrangian } torus fibration for general quintic {Calabi-Yau } hypersurfaces in \cite{lag3} and {Calabi-Yau } hypersurfaces in more general toric varieties in \cite{tor}.\\\\
\se{The Mirror construction and dual polyhedra}
It is well known that the mirror familly $\{Y_\psi\}$ of the family $\{X_\psi\}$ can be given by $Y_\psi=X_\psi/G$ with the induced complex structure and the {Calabi-Yau } orbifold metric. $G=<g_0, g_1, g_2, g_3>$ is the symmetry group of $X_\psi$. $G$ acts on ${\bf CP^4}$ as follows\\
\begin{eqnarray*}
g_0&=&(\xi, 1, 1, 1, \xi^{-1}),\\
g_1&=&(1, \xi, 1, 1, \xi^{-1}),\\
g_2&=&(1, 1, \xi, 1, \xi^{-1}),\\
g_3&=&(1, 1, 1, \xi, \xi^{-1}),
\end{eqnarray*}
where $\xi$ is a fifth root of unity. By analyzing the action of $G$ on $X_\psi$, it should be clear that $G\cong {\bf Z}_5^3$ will map every special Lagrangian fiber to itself, and is conjugate to the standard action of ${\bf Z}_5^3$ on $T^3$. The special Lagrangian fibration $f:\ X_\psi\rightarrow B$ will naturally induce special Lagrangian fibration $\hat{f}:\ Y_\psi\rightarrow B$. For any $p\in B$, $\hat{f}^{-1}(p)=f^{-1}(p)/G$, especially $\hat{f}^{-1}$ can be identified with $f^{-1}(p)$ quite canonically.\\\\
SYZ construction predict that the $T^3$ fibrations $f$ and $\hat{f}$ when restrict to $V_1\subset B$ should be dual to each other. But the above discussion implies that the two fibrations can be natually identified to each other instead of dual, especially the monodromy for $\hat{f}$ is the same as the monodromy for $f$. The key point is that the mirror map induce nontrivial identification $s:\ B\rightarrow B$ of base $B$ as defined in (\ref{ca}). Fibrations $\hat{f}$ and $B\circ f$ are actually dual to each other when restricted to $V_1\subset B$. This corresponds to the fact that the monodromy around $P_{ij}$ is dual to the monodromy around $P_{\overline{ij}}$. The reason why this is the correct interpretation is that the two fibrations should actually be written as\\
\[
f:\ X_\psi\longrightarrow \Delta,\ \ \ \hat{f}:\ Y_\psi\longrightarrow \hat{\Delta}
\]
where $\hat{\Delta}$ is the dual polyhedron of $\Delta$. The only canonical map from $\Delta$ to $\hat{\Delta}$ is the one corresponding to map $s$.\\\\
Now let us analyze the singular fibers of $\hat{f}$. For the type $I_5$ singular fiber $I_5\times S^1$, one ${\bf Z}_5$ will rotate $S^1$, another ${\bf Z}_5$ will rotate each $S^2$ in $I_5$ while keeping the nodal points fixed, the ${\bf Z}_5$ left will permute the five $S^2$'s in $I_5$. The quotient will natually be $I_1\times S^1$, we call it type $I$. A type $I$ fiber has Euler number $0$.\\\\
For the type $II_{5\times 5}$ singular fiber over $P_{ijk}$, two ${\bf Z}_5$'s will act on $T^2\cong T^3/{\gamma^m_l}$ as indicated by the two arrows in the following picture, which move a sixgon along the arrows. The ${\bf Z}_5$ left will act on $\gamma^m_l$ in standard way, which do not affect topology. Upon the action, edges of a fundamental reigon in the $T^2$ as indicated as darker area in the picture will be identified as indicated in the picture and turn into a $T^2$, with image of the graph $\Gamma_{ijk}$ turn into $\hat{\Gamma}_{ijk}$, topologically, a contractor of ``a pair of pants''. The resulting singular fiber can be identified as $S^1\times T^2$ with $S^1$'s over $\hat{\Gamma}_{ijk}$ collapsed to points. We will call it type $II$ singular fiber. A type $II$ fiber has Euler number $1$.\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=5in
\epsffile{six3c.eps}}
\end{center}
\caption{Type $II$ fiber}
\end{figure}\\
The quotients of the type $III_5$ singular fibers are very easy to describe. Two ${\bf Z}_5$'s will act on $T^2$ in standard way, topologically cause no change. The ${\bf Z}_5$ left will rotate among the chain of the five suspensions of $T^2$. The resulting singular fiber will simply be a suspension of $T^2$ with the two pole points identified as indicated in the following picture. We will call it type $III$ fiber. A type $III$ fiber has Euler number $-1$.\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=5in
\epsffile{p6.eps}}
\end{center}
\caption{Type $III$ fiber}
\end{figure}\\
It is interesting to see what is the image of the singular point sets\\
\[
\Sigma_{ijk}=\{[z]\in {\bf CP^4}| z_i^5 + z_j^5 + z_k^5 =0, z_l=0\ \ {\rm for}\ l\in\{1,2,3,4,5\}\backslash\{i,j,k\}\}
\]
under the quotient map. It is easy to see that $G$ map $\Sigma_{ijk}$ to itself. Two of the ${\bf Z}_5$'s act as identity, and the ${\bf Z}_5$ left acts freely on $\Sigma_{ijk}$. So $\Sigma_{ijk}$ is a 5-sheet cover of the quotient $\hat{\Sigma}_{ijk}$. Since\\
\[
5\cdot \chi(\hat{\Sigma}_{ijk})=\chi(\Sigma_{ijk})=-10
\]
we have\\
\[
\chi(\hat{\Sigma}_{ijk})=-2,\ \ g(\hat{\Sigma}_{ijk})=0.
\]
Namely $\hat{\Sigma}_{ijk}$ is a Riemann sphere. Their union is exactly the singular point set of the quotient $Y_{\psi}$.\\\\
Now we can try to compute the Leray sqectral sequence of the fibration $\hat{f}$. By analyzing the singular fibers described at above, we can esaily see that\\
\[
R^0 \hat{f}_*{\bf Q}\cong R^3 \hat{f}_*{\bf Q}\cong {\bf Q}
\]
and
\[
R^1 \hat{f}_*{\bf Q}\cong (R^2 \hat{f}_*{\bf Q})^\vee \cong i_* E_1
\]
These combine with computation from last section, give us\\
\begin{prop}
$E_2$ term of the Leray spectral sequence of fibration $\hat{f}$ is\\
\[
\begin{array}{cccc}
1 & 0 & 0 & 1\\
0 & 1 & 1 & 0\\
0 & 1 & 1 & 0\\
1 & 0 & 0 & 1
\end{array}
\]
\end{prop}
\begin{flushright} $\Box$ \end{flushright}
What follows is a discussion of the well known crepant resolution of $Y_{\psi}$. For completeness of mirror picture, we include it here. We need some degression on toric varieties. Let $X=Spec({\bf C}[C])$ be an affine toric variety, where
\[
C=<v^1, v^2, \cdots, v^n>_{\bf Q^+}\cap M \subset M=<e^1, e^2, \cdots, e^n>_{\bf Z}
\]
is a cone in lattice $M$. Following convention, let $N$ be the dual lattice of $M$ and ${\check C}$ be the dual cone of $C$. Suppose
\[
{\check C}=<v_1, v_2, \cdots, v_n>_{\bf Q^+}\cap N \subset N=<e_1, e_2, \cdots, e_n>_{\bf Z}.
\]
It is well known that a birational modification $X'$ of $X$ corresponds to subdividing ${\check C}$ into union of smaller cones (we require each new cone to be generated by $n$ elements), one dimensional edges of these cones correspond to toric divisors in $X'$. Let $D_v$ denote the toric divisor corresponding to the 1-dimensional cone generated by $v$. We will use $X_0$ to denote the smooth part of $X$. It is obvious that the singular part of $X$ is of codimension greater or equal to two. We need the following result.\\
\begin{prop}
There is a canonical nowhere vanishing holomorphic n-form $\Omega$ unique up to constant on $X_0$. Assume $X'$ is a birational modification of $X$, then $\Omega$ can be extend to $X_0$ as a nowhere vanishing holomorphic n-form if and only if the generater of any of the 1-dimension subcone coming from the subdivision is a convex combination of $ v_1, v_2, \cdots, v_n$.
\end{prop}
{\bf Proof:} It is well know that there is a canonical holomorphic n-form $\Omega_0$ on the $n$-torus defined as
\[
\Omega_0=\frac{{\rm d}e^1}{e^1}\wedge\frac{{\rm d}e^2}{e^2}\wedge\cdots \wedge\frac{{\rm d}e^n}{e^n}
\]
For any primary $v\in N$, we can get $D_v=Spec({\bf C}[v^\bot])$. $\Omega_0$ will have pole of order 1 around $D_v$. To cancel the pole, we need to multiply a function, which vanish to order one at $D_v$. These kind of function will be in\\
\[
v^\vee =\{w\in M|<v, w>=1\}
\]
Suppose that\\
\[
\bigcap _{i=1}^n v_i^\vee =\{w\}
\]
then $\Omega=w\cdot\Omega_0$ is exactly what we need.\\\\
Suppose that $v$ is the generater of an 1-dimension subcone coming from the subdivision corresponding to a birational modification $X'$. Then in order to extend $\Omega=w\cdot\Omega_0$ to $D_v\subset X'$, $v$ must satisfy $<v, w>=1$, which is the same as that $v$ is a convex combination of $ v_1, v_2, \cdots, v_n$.
\begin{flushright} $\Box$ \end{flushright}
{\bf Remark:} The above proposition must be quite well known in toric geometry. But we do not know exactly where it was located.\\\\
$Y_{\psi}$ is a singular {Calabi-Yau } manifold. It has singularity at $\tilde{P}_{ij}$ and $\hat{\Sigma}_{ijk}$, where $\tilde{P}_{ij}$ is the singular point of the fiber over $P_{ij}$. Around $\tilde{P}_{ij}$, $Y_{\psi}$ has quotient singularity. It can be modeled by affine toric variety $X=Spec({\bf C}[C])$, where\\
\[
C=<5e^1, 5e^2, 5e^3>_{\bf Q^+}\cap M \subset M=<5e^1, 5e^2, 5e^3, \sum_{i=1}^3 e^i>_{\bf Z}
\]
Here $\{e^i\}$ denote the standard base of ${\bf Z}^3$. We would like to get a crepent resolution of the toric variety $X=Spec({\bf C}[C])$, for which we have to subdivide the dual cone\\
\[
{\check C}=<e_1, e_2, e_3>_{\bf Q^+}\cap N \subset N=M^\vee,
\]
here $\{e_i\}$ denote the standard base of the dual ${\bf Z}^3$. By above discussion, we only need to look at the plane passing through the three points $e_1, e_2, e_3$. ${\check C}$ will cut off a triangle with this three points as vertices in the plane. All the lattice points in $N$ that is in this triangle are\\
\[
v_{ijk}=\frac{i}{5}\cdot e_1+\frac{j}{5}\cdot e_2 +\frac{k}{5}\cdot e_3,\ \ i,j,k\geq 0\ \ i+j+k=5
\]
We will divide this triangle as indicated in following picture. This division will result in a subdivision of ${\check C}$ which gives us a desired crepent resolution.\\
\begin{figure}[h]
\begin{center}
\leavevmode
\hbox{%
\epsfxsize=5in
\epsffile{tri.eps}}
\end{center}
\caption{}
\end{figure}\\
There are exactly three curves of singularity $\hat{\Sigma}_{ijk}$ for $k\in \overline{\{ij\}}$ that pass through the singular point $\tilde{P}_{ij}$. In the picture, they correspond to the interiors of the three edges of the first triangle, while the singular point correspond to the interior of the triangle. After the birational modification, we will have newly added exceptional divisors that correspond to dots other than $e_1,e_2,e_3$. The ones in the interiors of the three edges correspond to exceptional divisors coming from the three curves of singularity. The ones in the interior of the first triangle correspond to exceptional divisors coming from the singular point. So each curve of singularity will give rise to 4 exceptional divisors and each singular point will give rise to 6 exceptional divisors. There are 10 curves of singularity and 10 singular points in $Y_\psi$. So all together we will have $4\times 10 + 6\times 10= 100$ of exceptional divisors in the smooth {Calabi-Yau } $\tilde{Y}_\psi$ we get as the crepant resolution of $Y_\psi$. This verify the fact\\
\begin{prop}
\[
h^0(\tilde{Y}_\psi)=h^6(\tilde{Y}_\psi)=1, \ h^1(\tilde{Y}_\psi)=h^5(\tilde{Y}_\psi)=0,\ h^2(\tilde{Y}_\psi)=h^4(\tilde{Y}_\psi)=101,\ h^3(\tilde{Y}_\psi)=4.
\]
\end{prop}
\begin{flushright} $\Box$ \end{flushright}
{\bf Acknowledgement:} I would like to thank Qin Jing for many very stimulating discussions during the course of my work, and helpful suggestions while carefully reading my early draft. I would also like to thank Prof. Yau for his constant encouragement.\\\\
|
\section*{Tables}
\begin{table}[hbt]
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|c|} \hline
$f$ & $N$ & $\delta t/\tau$ & $t_{max}/\delta t$ & $R_{G}/\sigma_{LJ}$ &
$(R_{c}/\sigma_{LJ}) + 1/2$ & $R_{d}/\sigma_{LJ}$ & $\lambda$\\ \hline
5 & 100 & 0.004 & 500000 & 13.53 & 0.65 & 1.39 & 0.61\\
10 & 50 & 0.003 & 400000 & 10.37 & 1.1 & 1.21 & 0.66\\
10 & 100 & 0.003 & 400000 & 16.18 & 1.1 & 0.89 & 0.64\\
10 & 150 & 0.003 & 400000 & 19.71 & 1.1 & 1.31 & 0.60\\
10 & 200 & 0.003 & 400000 & 24.52 & 1.1 & 1.42 & 0.67\\
18 & 50 & 0.002 & 350000 & 11.19 & 1.25 & 1.38 & 0.68\\
18 & 100 & 0.002 & 350000 & 17.10 & 1.25 & 1.64 & 0.65\\
30 & 50 & 0.002 & 350000 & 12.22 & 1.6 & 1.89 & 0.66\\
50 & 50 & 0.002 & 350000 & 13.35 & 1.8 & 2.40 & 0.69\\ \hline
\end{tabular}
\end{center}
\caption{ \label{parameters.table}
List of the simulation parameters and the
corresponding results for
$R_{G}$ and $\lambda=\sigma/2R_{G}$.
}
\end{table}
\section*{Figure captions}
\begin{figure}[hbt]
\caption{ \label{snapshot.plot}
Typical configuration for two stars with $f=10$ and
$N=50$. The distance between the central core particles, which
are shown as big black spheres, is $r=5.2\sigma_{LJ}$.
The gray and light gray monomers are belonging to the
first and second star respectively.
}
\end{figure}
\begin{figure}[hbt]
\caption{ \label{invforce.plot}
Reduced inverse force $k_{B}T/(FR_{G})$
between the centers of two star polymers
(for $f=10$ and $N=50$) versus reduced distance $r/R_{G}$.
The error bars were obtained by
averaging over the results of $8$ independent
simulations.
}
\end{figure}
\begin{figure}[hbt]
\caption{ \label{results.plot1}
Simulation results (symbols) and theoretical results (lines)
for the reduced effective force $FR_{G}/k_{B}T$ versus
reduced distance $(r-2R_{d})/R_{G}$.
a) for $f=5, 10, 18$ and $N=100$,
b) for $f=18, 30, 50$ and $N=50$.
}
\end{figure}
\begin{figure}[hbt]
\caption{ \label{lnforce.plot}
Logarithm of the reduced force $\ln\left(FR_{G}/k_{B}T\right)$
versus reduced distance $(r-2R_{d})/R_{G}$ for $f=10$ and $N=50$. The error bars were obtained by
averaging over the results of $8$ independent
simulations.
}
\end{figure}
\vfill
\end{document}
|
\section{Introduction}
A few years ago we have proposed
a quark-parton model of cumulative phenomena
in the interactions with nuclei \cite{NPB94,YF97}
based on
perturbative QCD calculations of the corresponding quark
diagrams near the thresholds, at which other quarks
("donors") in the nuclear flucton transfer all their
longitudinal momenta to the distinguished active quark
and become soft.
Consider the scattering of
a hadronic projectile off a nucleus with the momentum $P$
in the c.m. system.
At high energies
the momentum $K$ of the produced pion belongs to
the cumulative region if $K_z>P_z/A$.
As a reasonable first approximation, we treat the nucleus as a
collection of $N=3A$ valence quarks, which, on the average, carry each
longitudinal momentum $x_{0}P_{z}/A$ with $x_{0}=1/3$.
In our approach the cumulative pion production proceeds in two steps.
First a valence quark with a scaling variable $x>1$ is created.
Afterwards
it decays into the observed hadron with its scaling variable $x$ close
to the initial cumulative quark's one. This second step is described by
the well-known quark fragmentation functions
\cite{CapellaT81} and will not be
discussed here.
The produced cumulative ("active") quark acquires
the momentum much greater than $x_{0}P_{z}/A$ only if this
quark has interacted by means of gluon exchanges
with other $p$ quarks of flucton ("donors")
and has taken some of their longitudinal momenta (see Fig.1).
If this active quark accumulates all
longitudinal momentum of these $p$ quarks then
$K_z=(p+1)x_0P_z/A$ and the donors become soft.
It is well-known that interactions which make the longitudinal momentum
of one of the quark equal to zero may be treated by perturbation theory
\cite{Brodsky92}. This allows to calculate the part of Fig. 1
responsible for the creation of a cumulative quark explicitly.
This was done in \cite{NPB94,YF97}, to which papers we refer the
reader for all the details. As a result we were able to explain the
exponential fall-off of the production rate in the cumulative region.
Since with the rise of $x$ the active quark has to interact with
a greater number of donors, one expects that its average transverse momentum
also grows with $x$. Roughly one expects that $\langle K_{\perp}^2 \rangle$
is proportional to the number of interactions, that is, to $x$.
In \cite{NPB94,YF97}
this point was not studied:
we have limited ourselves with the inclusive cross-section
integrated over the transverse momenta,
which lead to some simplifications.
The aim of the present paper is to find
the pion production rate dependence on
the transverse momentum
and the mean value of the latter
as a function of $x$ in the cumulative region.
This dependence and also the magnitude of $\langle K_{\perp}^2 \rangle$
have been studied experimentally. The comparison of our predictions
with the data allows to obtain further support for our model and
fix one of the two its parameters (the infrared cutoff).
\section{The $K_{\perp}$ dependence}
Repeating the calculations of the diagram in Fig.1 described in
\cite{NPB94,YF97} but not
limiting ourselves with the inclusive cross-section
integrated over the transverse momentum,
we readily find that all dependence upon
the transverse momentum $K_{\perp}$ of the produced particle
is concentrated in a factor:
\begin{equation}
J (K_{\perp})=
\int
\rho_A(\underbrace{r,...,r}_{p+1}|\underbrace{\ol{r},...,\ol{r}}_{p+1})
G(c_1,...,c_p)
\prod_{i=1}^{p}
\lambda(c_i-r)\lambda(c_i-\ol{r})d^2 c_i
e^{i(\ol{r}-r)K_{\perp}}
d^2 r d^2 \ol{r}
\end{equation}
Here $\rho_A$ is the (translationally invariant)
quark density matrix of the nucleus:
\begin{equation}
\rho_A(r_i|\ol{r}_i)\equiv
\int
\psi_{\perp A}(r_i,r_m)
\psi^*_{\perp A}(\ol{r}_i,r_m)
\prod_{m=p+2}^{N} d^2 r_m
\end{equation}
where
$\psi_{\perp A}$ is the transverse part of
the nuclear quark wave function. The propgation of soft donor quarks
is decribed by
\begin{equation}
\lambda(c)=\frac{K_0(m|c|)}{2\pi}
\label{lam}
\end{equation}
where $m$ is the constituent quark mass and $K_0$
is the modified Bessel function (the Mac-Donald function).
The interaction with the projectile contributes a factor
\begin{equation}
G(c_1,...,c_p)=
\int
\prod_{i=1}^{p}
\sigma_{qq}(c_i-b_i)
\eta_H(b_1,...,b_p) d^2 b_i
\end{equation}
where $\sigma_{qq}(c)$ is the quark-quark cross-section
at a given value of impact parameter $c$ and
\begin{equation}
\eta_H(b_1,...,b_p)=
\sum_{L\geq p}\frac{L!}{(L-p)!}
\int|\psi_{\perp H}(b_i)|^2
\delta^{(2)}(\frac{1}{L}\sum_{i=1}^{L} b_i)
\ d^2 b_{p+1}...d^2 b_L
\end{equation}
is a multiparton distribution in the projectile, expressed via
the transverse part of
its partonic wave function $\psi_{\perp H}$ .
If we integrate $J (K_{\perp})$ over $K_{\perp}$ we come back
to our old result (Eq. (33) in \cite{NPB94}):
$$
\int J (K_{\perp})
\frac{d^2 K_{\perp}}{(2\pi)^2}=
\rho_A(\underbrace{0,...,0}_{p+1}|\underbrace{0,...,0}_{p+1})
\int
G(c_1,...,c_p)
\prod_{i=1}^{p} \lambda^2(c_i-r)
d^2 c_i d^2 r
$$
If one assumes factorization of the multiparton distribution
$\eta_H(b_1,...,b_p)$ then\\
$G(c_1,...,c_p)$ also factorizes:
\begin{equation}
G(c_1,...,c_p)=
\prod_{i=1}^{p}
G_0(c_i)
\label{facG}
\end{equation}
Following \cite{YF97} we use the quasi-eikonal approximation
for $\eta_H$:
$$
\eta_H(b_1,...,b_p)=
\xi^{(p-1)/2}\nu_H^{p}\prod_{i=1}^{p}\eta_H(b_i)
$$
where $\xi$ is the quasi-eikonal diffraction factor,
$\nu_H^{}$ is the mean number of partons in the projectile hadron
and the single parton distribution $\eta_H(b)$
is normalized to unity.
In a Gaussian approximation
for $\sigma(c)$ and $\eta_H(b)$ we find:
$$
G_0(c)=
\xi^{\frac{1}{2}-\frac{1}{2p}}
\frac{\nu_H\sigma_{qq}}{\pi r_{0H}^2}
e^{-\frac{c^2}{r_{0H}^2}}
$$
where $\sigma_{qq}$ is the total quark-quark cross-section,
$r_{0H}^2=r_0^2+r_H^2$, $r_0$ and $r_H$ are the widths of
$\sigma(c)$ and $\eta_H(b)$ respectively.
With the factorised $G(c_1,...,c_p)$ (\ref{facG}) we have
$$
J (K_{\perp})=
\int
\rho_A(\underbrace{0,...,0}_{p+1}|\underbrace{\ol{r}-r,...,\ol{r}-r}_{p+1})
j^p(r,\ol{r})
e^{i(\ol{r}-r)K_{\perp}}
d^2 r d^2 \ol{r}
$$
where
$$
j(r,\ol{r})=
\int d^2 c
G_0(c)
\lambda(c-r)
\lambda(c-\ol{r})
$$
We also have used the translational invariance of the $\rho$-matrix.
Note that near the real threshold we have no spectators and
$$
\rho_A(\underbrace{0,...,0}_{p+1}|\underbrace{\ol{r}-r,...,\ol{r}-r}_{p+1})
=\rho_A(\underbrace{0,...,0}_{p+1}|\underbrace{0,...,0}_{p+1})
$$
In any case large $K_{\perp}$ corresponds to small $\ol{r}-r$ so
we factor $\rho_A$ out of the integral in zero point.
In the rest integral we pass to the variables
$$
B=\frac{r+\ol{r}}{2}, \hs 1 b=\ol{r}-r
$$
and shift the integration variable $c$, then
\begin{equation}
J (K_{\perp})=
\rho_A(\underbrace{0,...,0}_{p+1}|\underbrace{0,...,0}_{p+1})
\int
j^p(B,b)
e^{ibK_{\perp}}
d^2 b d^2 B
\end{equation}
where
\begin{equation}
j(B,b)=
\int
G_0(B+c)
\lambda(\frac{b}{2}-c)
\lambda(\frac{b}{2}+c)
d^2 c
\label{j}
\end{equation}
\section{The calculation of $\langle |K_{\perp}|\rangle $}
Now we would like to find the width of the distribution on $K_{\perp}$
as a function of $p$ or what is the same of the cumulative
number $x=(p+1)/3$.
From the mathematical point of view it is simpler to calculate
the mean squared width of the distribution $\langle K_{\perp}^2\rangle $.
Unfortunately in our case this quantity is logarithmically
divergent at large $K_{\perp}$.
This divergency results from the behavior of $j(B,b)$
at small $b$. This behavior is determined by the behavior of the
$ \lambda(b)=K_0(m|b|)/(2\pi) $ (\ref{lam}),
which has a logarithmical singularity
at $|b|=0$. Smooth $G_0(B+c)$ does not affect this behavior.
For this reason we shall rather calculate $\langle |K_{\perp}|\rangle $:
\begin{equation}
\langle |K_{\perp}|\rangle
=\frac{1}{J_N}
\int
j^p(B,b)
|K_{\perp}|e^{ibK_{\perp}}
d^2 b d^2 B
\frac{d^2 K_{\perp}}{(2\pi)^2}
\end{equation}
where $J_N$ is the same integral as in the numerator
but without $|K_{\perp}|$.
Presenting $|K_{\perp}|$ as $K_{\perp}^2/|K_{\perp}|$ and
$K_{\perp}^2$ as the Laplacian $\Delta_b$ applied to
the exponent we find
$$
\langle |K_{\perp}|\rangle
=-\frac{1}{J_N}
\int
j^p(B,b)
\Delta_b e^{ibK_{\perp}}
d^2 b d^2 B
\frac{d^2 K_{\perp}}{|K_{\perp}|(2\pi)^2}
$$
Twice integrating by parts
and using the formula
$$
\int
\frac{d^2 K_{\perp}}{|K_{\perp}|}e^{ibK_{\perp}} =
\frac{2\pi}{|b|}
$$
we find
$$
\langle |K_{\perp}|\rangle
=-\frac{1}{2\pi J_N}
\int
\frac{1}{|b|}
\Delta_b j^p(B,b)
d^2 b d^2 B
$$
Now we again integrate by parts once to find
$$
\langle |K_{\perp}|\rangle
=-\frac{1}{2\pi J_N}
\int
d^2 B
\frac{d^2 b}{|b|^2}
(n_b \nabla_b) j^p(B,b)
$$
where $n_b=b/|b|$. This leads to our final formula
\begin{equation}
\langle |K_{\perp}|\rangle
=-\frac{p}{2\pi J_N}
\int
d^2 B
\frac{d^2 b}{|b|^2}
j^{p-1}(B,b)
(n_b \nabla_b) j(B,b)
\label{Kperp}
\end{equation}
where $j(B,b)$ is given by (\ref{j}),
$ \lambda(b) $ is given by (\ref{lam}) and
$$
J_N=\int d^2 B j^{p}(B,b=0)
$$
\section{Approximations}
To simplify numerical calculations we make some additional
approximations, which are not very essential but are well supported
by the comparison with exact calulations at a few sample points.
As follows from the
the asymptotics of $K_0(z)$ at large $z$
$$
K_0(z)\simeq\sqrt{\frac{\pi}{2z}}e^{-z}
$$
the width of $\lambda(b)$ (\ref{lam}) is of the order $m^{-1}$.
The function $G_0$ is smooth in the vicinity of the origin
and its width $r_{0H}=\sqrt{r_0^2+r_H^2}$ is substancially
larger than the width of $\lambda$.
For this reason we factor $G_0(B+c)$ out of the integral (\ref{j})
over $c$ at the point $B$:
\begin{equation}
j(B,b)=
G_0(B)\Lambda(b),
\hs 1
\Lambda(b) \equiv
\int
\lambda(c)
\lambda(c-b)
d^2 c=
\frac{|b|}{4\pi m}K_1(m|b|)
\label{Lam}
\end{equation}
Then we find that the integrals over $B$ and $b$ decouple
\begin{equation}
J (K_{\perp})=
\rho_A(\underbrace{0,...,0}_{p+1}|\underbrace{0,...,0}_{p+1})
\int G^p_0(B) d^2 B
\int \Lambda^p(b)
e^{ibK_{\perp}}
d^2 b
\end{equation}
In this approximation we find that $\langle |K_{\perp}|\rangle $
depends only
on one parameter - the constituent quark mass $m$,
which in our approach plays the role of an infrared cutoff:
\begin{equation}
\langle |K_{\perp}|\rangle
=pm
\int_0^{\infty} dz K_0(z) (zK_1(z))^{p-1}
\label{apprKperp}
\end{equation}
This allows to relate $m$ directly to the experimental data on
the transverse momentum dependence.
\section{Comparison with the data and discussion}
The integral in (\ref{apprKperp}) can be easy calculated
numerically. For values of $p=1,...,12$ it is very well
approximated by a power dependence (see Fig.2), so that we obtain
\begin{equation}
\langle |K_{\perp}|\rangle/m
=1.594\, p^{0.625}
\end{equation}
As we observe, the rise of $\langle |K_{\perp}|\rangle$ turns out to be
even faster than
expected on naive physical grounds mentioned in the Introduction
($\sim\sqrt{p}$).
The resulting plots for $\langle |K_{\perp}|\rangle^2$
as a function of the cumulative number $x=(p+1)/3$
at different values of parameter $m$ are shown in Fig.3
together with avaiable experimental data from \cite{Boyarinov94}
on $\langle K_{\perp}^2\rangle$ for pion production obtained in
experiments
\cite{Boyarinov94}-\cite{Boyarinov87}
with 10 $GeV$ protons
and \cite{Baldin82, Baldin83} with 8.94 $GeV$ protons.
Note that earlier publications of the first group
\cite{Boyarinov92,Boyarinov87}
reported a much stronger increase of $\langle K_{\perp}^2\rangle$
with $x$, up to value 2 $(GeV/c)^2$ at $x=3$ for pion production.
In our approach such an increase would require the quark mass to be as high
as $m \simeq 225 MeV$.
In a more recent publication \cite{Boyarinov94}
the rise of $\langle K_{\perp}^2\rangle$ is substancially weaker
(it corresponds to $m \simeq 175 MeV$ in our approach).
The authors of \cite{Boyarinov94} explain this
by new experimental data obtained and by a cutoff $K_{\perp max}$
introduced in calculations of $\langle K_{\perp}^2\rangle$ in \cite{Boyarinov94}.
The introduction of this cutoff considerably
(approximatly two times)
decreases the experimental value of $\langle K_{\perp}^2\rangle$ at $x=3$.
In our opinion this is a confirmation that the
cumulative pion production rate only weakly decreases with $K_{\perp}$
in the cumulative region so that the
the integral over $K_{\perp}^2$ which enters the definition of
$\langle K_{\perp}^2\rangle$ is weakly convergent or even divergent,
as in our approach.
Undoubtedly presentation of the experimental data in terms of
the mean value
$\langle |K_{\perp}| \rangle^2$,
rather than $\langle K_{\perp}^2\rangle$ should reduce
the dependence on the cutoff
$K_{\perp max}$ and make the results more informative.
One of the ideas behind the investigations of the cumulative
phenomena is that they may be a manifestation
of a cold quark-gluon plasma formed when several nucleons overlap in the
nuclear matter. In \cite{NPB94} we pointed out that our model does not
correspond to this picture. It implies coherent interactions of the
active quark with donors and, as a result, strong correlations between the
longitudinal and transverse motion. Predictions for the dependence of
$\langle |K_{\perp}| \rangle$ on $x$ are also different. From the cold
quark-gluon plasma model one expects $\langle |K_{\perp}| \rangle$ to behave
as $x^{1/3}$, since the Fermi momentum of the quarks inside the overlap
volume is proportional to the cubic root of the quark density. Our model
predicts a much faster increase, with a power twice larger. The
experimental data seem to support our predictions.
\section{Acknowledgments}
The authors are greatly thankful to Prof. P.Hoyer who attracted
their attention to the problem.
This work is supported by the Russian Foundation for Fundamental
Research, Grant No. 97-02-18123.
\newpage
|
\section{Introduction}
When the core of a large star ($M \ge 8 M_{\odot}$) runs out of
nuclear fuel, it collapses and forms a proto-neutron star. The total energy
released in the collapse, i.e., the gravitational binding energy of
the core ($E_B \sim G_N M_ {\odot}^2/R$ with $R \sim$ 10 km), is about
$3 \times 10^{53}$ ergs; $\sim$ 99\% of that is carried away by
neutrinos and antineutrinos, the particles with the longest mean free
path. It is believed that neutrinos of all three flavors
are emitted with approximately equal luminosities
over a timescale of several seconds.
Those flavors which interact the most with the matter will decouple at
the largest radius and thus the lowest temperature. The $\nu_\mu$ and
$\nu_\tau$ neutrinos and their antiparticles
have only neutral-current interactions with
the matter, and therefore leave with the highest temperature, about 8
MeV (or $\langle E \rangle \simeq$ 25 MeV). The $\bar{\nu}_e$ and
$\nu_e$ neutrinos have also charged-current interactions, and so leave
with lower temperatures, about 5 MeV ($\langle E \rangle \simeq$ 16
MeV) and 3.5 MeV ($\langle E \rangle \simeq$ 11 MeV), respectively.
The $\nu_e$ temperature is lower because the material is neutron-rich
and thus the $\nu_e$ interact more than the $\bar{\nu}_e$. The
observation of supernova
$\nu_\mu$ and $\nu_\tau$ neutrinos and their antiparticles
would allow the details of the picture above to be tested.
In this talk I concentrate on two aspects of the neutrino signal:
\begin{itemize}
\item The possibility of measuring or constraining the mass of the
$\nu_\tau$ and/or $\nu_\mu$.
\item The possibility of locating the supernova by its neutrino signal,
independently of or prior to the optical observation.
\end{itemize}
The physics of these task is straightforward, but there are complications
due to:
\begin{itemize}
\item The finite statistics of the neutrino signal.
\item The finite time duration of the signal.
\end{itemize}
The details of the work reported here can be found in the joint
work with John Beacom of Caltech \cite{SK,SNO,point}.
One can find a much more complete list
of the relevant earlier references there.
Numerical supernova models suggest that the neutrino luminosity rises
quickly over a time of order 0.1 s, and then falls over a time of
order several seconds. The rise is so fast
that the details of its shape are largely irrelevant
for our task. We model
the luminosity fall by an
exponential with time constant $\tau$ = 3 s. The luminosity
then has a width of about 10 s, consistent with the SN 1987A
observations. Later, I will show how the conclusions depend
on the luminosity decay time constant $\tau$.
\section{Neutrino mass determination}
The requirement that neutrinos do not overclose the universe gives a
bound for the sum of masses of stable neutrinos (see e.g., \cite{Raffelt}):
\begin{equation}
\sum_{i=1}^3 m_{\nu_i} \le 100 {\rm\ eV}\,.
\label{eq:cosmo}
\end{equation}
However, laboratory kinematic tests of neutrino mass currently give limits
for the masses compatible with the above cosmological bound only for
the electron neutrino, $m_{\bar{\nu}_e} \le 5$ eV \cite{Belesev}.
For the $\nu_\mu$ and $\nu_\tau$ they far
exceed the cosmological bound: $m_{\nu_\mu} < 170$ keV\cite{RPP}, and
$m_{\nu_\tau} < 18$ MeV \cite{RPP}. It is very unlikely that
these mass limits can improve by the necessary orders
of magnitude any time soon.
When neutrinos are emitted by a supernova, even a tiny
mass will make the velocity less than for a massless particle,
and will cause a
measurable delay in the arrival time. A neutrino with a mass $m$ (in
eV) and energy $E$ (in MeV) will experience an energy-dependent delay
(in s) relative to a massless neutrino in traveling over a distance D
(in 10 kpc) of
\begin{equation}
\Delta t(E) = 0.515 \left(\frac{m}{E}\right)^2 D\,.
\label{eq:delay}
\end{equation}
For a supernova at 10 kpc distance (approximately at the
center of the galaxy), the delay for $\nu_e$ and $\bar{\nu}_e$
will be negligible, and their signal can
be used as a reference clock. The $\nu_\mu$ and $\nu_\tau$ neutrinos and
their antiparticles will interact only by the neutral current. Thus,
in order to determine the $\nu_\tau$ and/or $\nu_\mu$ mass, we should find
ways of separating the neutral and charged current signals,
and of determining the possible time delay of the former
with respect to the latter.
There are three neutral current reactions that give rise to
potentially measurable signals:
a) neutrino-electron scattering (we will show below
that it is difficult to separate
the charged and neutral current events in that case),
b) neutral current excitation
of $^{16}$O nuclei in water, followed by the $\gamma$
emission as suggested in \cite{LVK},
and c) the neutral current deuteron disintegration (relevant for SNO).
In Table \ref{tab:rate} I show the corresponding numbers of events
(see \cite{SK,SNO} for details how the table
was made and refs. \cite{SKD,SNOD} for
description of the detectors) for the individual
reactions, calculated for the ``standard'' supernova defined above.
\begin{table}[h]
\caption{Calculated numbers of events expected in SK and SNO.
In SNO events in 1 kton of D$_2$O and in 1.4 kton of H$_2$O are added.
By $\nu_x$ we denote the combined effect of $\nu_\mu$ and $\nu_\tau$,
each accounts for half of the events.
In all except the top row, the events caused by $\nu$ and $\bar{\nu}$
are added.}
\label{tab:rate}
\vspace{5 mm}
\begin{center}
\small
\begin{tabular}{|l|l|l|}
\hline
reaction & events in SK & events in SNO \\
\hline\hline
$\bar{\nu}_e + p \rightarrow e^+ + n$ & 8300 & 365 \\
\hline
$\nu_e + d \rightarrow e^- + p + p $ & - & 160 \\
$\bar{\nu}_e + d \rightarrow e^+ + n + n $ & & \\
\hline
$\nu_x + d \rightarrow \nu_x + n + p $ & - & 400 \\
\hline
$\nu_x+ ^{16}{\rm O} \rightarrow \nu_x + \gamma + X$ & 710 & 50 \\
\hline
$\nu_x+ ^{16}{\rm O} \rightarrow \nu_x + n + ^{15}{\rm O} $ & - & 15 \\
\hline
$\nu_e + e^- \rightarrow \nu_e + e^-$ & 200 & 15 \\
\hline
$\nu_x + e^- \rightarrow \nu_x + e^-$ & 120 & 10 \\
\hline \hline
\end{tabular}
\end{center}
\end{table}
Given the assumed known time dependence of the supernova luminosity $L(t)$,
and assuming that all flavors develop in time the same way and keep their
temperatures constant, the arrival time of massive neutrinos is then described
by $L(t - \Delta t(E_{\nu}))$. Since in the neutral current scattering one cannot
determine the incoming neutrino energy $E_{\nu}$, we have at our disposal
only the {\it time distribution} of the events
\begin{equation}
\frac{dN}{dt} = C
\int dE_{\nu} f(E_{\nu}) \sigma(E_{\nu}) L(t - \Delta t(E_{\nu})) \,,
\label{eq:rate}
\end{equation}
where
$C$ is a constant proportional to $1/(D^2 \times \langle E_{\nu} \rangle)$
and $f(E_{\nu})$ is the thermal neutrino spectrum.
The only way one can decide whether there is
a time delay or not is to compare the
neutral current time distribution (called ``Signal'')
with the time distribution
of the charged current events (called ``Reference''). Since
$\nu_\tau$ and $\nu_{\mu}$ neutrinos and their antiparticles have higher
energies than $\nu_e$ and $\bar{\nu}_e$, the neutral current events will
contain a substantial fraction of possibly delayed events, while the charged
current events will have no delay.
It turns out, see \cite{SK}, that the most efficient way
to accomplish this is also the
simplest one, i.e., to use the diffence in the
{\it mean arrival time}:
\begin{equation}
\langle t \rangle_S = \sum_k t_k /N_S \,,
~ \langle t \rangle_R = \sum_k t_k /N_R \,,
\end{equation}
where $N_S (N_R)$ is the total number
of the Signal (Reference) events, and $t_k$
are the arrival times of the individual events.
The signature of neutrino mass is then the inequality
\begin{equation}
\langle t \rangle_S > \langle t \rangle_R \,,
\end{equation}
valid with significance beyond statistical fluctuations.
\begin{figure}[h]
\vspace*{13pt}
\begin{center}
\mbox{\epsfig{figure=moments1m3s.eps,width=8cm}}
\vspace*{13pt}
\fcaption{The results of the $\langle t \rangle$ analysis for a massive
$\nu_\tau$ in SK using the
$\gamma$ following $^{16}$O excitation.
In the upper panel, the relative frequencies of various
$\langle t \rangle_S - \langle t \rangle_R$ values are shown for a few
example masses. In the lower panel, the range of masses corresponding
to a given $\langle t \rangle_S - \langle t \rangle_R$ is shown. The
solid line is the 50\% confidence level, and the upper and lower
dashed lines are the 10\% and 90\% confidence levels, respectively.}
\label{fig:mom}
\end{center}
\end{figure}
The analysis below is based on the assumption that only one of the
neutrino flavors is massive, say $\nu_{\tau}$, and the other one,
$\nu_{\mu}$ in this case, is either massless or has so much smaller mass
that the corresponding time delay is negligible. The ``Signal'' then consist
of part that is delayed and another part that is not because it is either
caused by the massless $\nu_{\mu}$ or belongs to background that cannot
be separated from the signal since it has the same energy and angle, etc.
With these assumptions, the neutrino-electron scattering signal in SK
will contain 60 delayed events and 700 background events, since a rather large
number of the charged current $\bar{\nu}_e + p \rightarrow e^+ + n$ events
will be present in the forward cone. For the $\gamma$ signal from the
$^{16}$O excitation, the ratio delayed/background is a more favorable 355/885.
And in SNO the neutral current deuteron disintegration, with a single
neutron and no charged lepton, is characterized by the ratio 219/316.
The last two ratios above also show that events which look like
the neutral current (i.e., the true neutral current plus background with
similar characteristics) are
dominated by the response to $\nu_{\tau}$ and $\nu_{\mu}$ neutrinos.
Indeed, since the cross sections are the same for these two flavors,
one can simply multiply the numerators by a factor of two and make
the corresponding adjustment in the denominator, to obtain
the fraction of events caused by the $\nu_{\tau}$ and $\nu_{\mu}$ neutrinos.
So, for the $\gamma$ signal from the $^{16}$O excitation, that contribution
is about 57\%, and for the deuteron disintegration it is about 82\%. On the other
hand, for the neutrino-electron scattering it is only about 16\%.
By measuring the total number of the Signal events one can determine
the temperatures of the $\nu_{\tau}$ and $\nu_{\mu}$ neutrinos
(see \cite{SK,SNO}) with reasonable accuracy.
To judge the statistical significance of the delay, we used the Monte Carlo
simulation of a large number of supernovae for each mass value. We then
histogram the differences
$ \langle t \rangle_S - \langle t \rangle_R$, and find
the 10\%, 50\%, and 90\% confidence levels. Representative cases are
plotted in Fig. 1 for the $\gamma$ from $^{16}$O
excitation in SK. In the upper panel one can see that if,
e.g., $m_{\nu}$ = 75 eV
the most probable difference in average arrival times is about 0.2 s, and the
10 - 90 \% CL band is 0.1 - 0.3 s, clearly separated from the massless case.
In fact, the smallest recognizable mass is about 50 eV.
In SNO, using the deuteron disintegration,
the smallest recognizable mass is even
smaller, about 30 eV.
How does the mass sensitivity depend on the assumptions we made above?
Much of the analysis can be made analytically. We shall concentrate on
the dependence on the assumed temperatures $T$, distance $D$, and the
constant which characterizes the time duration
of the neutrino signal, $\tau$. The time delay
and its error depend on
\begin{equation}
\langle t \rangle_S - \langle t \rangle_R \sim (m/T)^2 D ~;~
\delta( \langle t \rangle_S - \langle t \rangle_R) \sim \tau D/\sqrt{T} ~.
\end{equation}
Since the significance of the result, and thus the smallest recognizable mass
$m_{lim}$,
is the ratio of these two quantities, we conclude that this neutrino mass limit
is, remarkably, independent on the distance $D$, and
\begin{equation}
m_{lim} \sim \sqrt{\tau} T^{3/4} ~.
\end{equation}
Clearly, the shorter the duration of the neutrino pulse
(i.e., smaller $\tau$), the
better the ability to determine the neutrino mass.
(We verified that detailed numerical simulation closely follows the
$\sqrt{\tau}$ scaling above)
Also, naturally, if e.g. the
$\nu_{\mu}$ and $\nu_{\tau}$ masses are close to each other, the mass limit
is improved, roughly by $\sqrt{2}$.
\section{ Supernova localization with neutrinos}
A future core-collapse supernova in our Galaxy will be detected by several
neutrino detectors around the world. The neutrinos escape from the
supernova core over several seconds from the time of collapse, unlike
the electromagnetic radiation, emitted from the envelope, which is
delayed by a time of order hours. In addition, the electromagnetic
radiation can be obscured by dust in the intervening interstellar
space. The question therefore arises whether a supernova can be
located by its neutrinos alone. The early warning of a supernova and
its location might allow greatly improved astronomical observations.
There are two types of techniques to locate a supernova by its
neutrinos. The first one is based on angular
distributions of the neutrino reaction products, which can be
correlated with the neutrino direction. In this case, a single
experiment can independently announce a direction and its error.
However, to suppress false alarms one
can demand coincidence with other experiments.
The second method of supernova
location is based on triangulation using two or more widely-separated
detectors. This technique would require significant and immediate
data sharing among the different experiments.
The theme of this section (for more details and more
complete reference list, see \cite{point})
is a careful and realistic assessment of
this question, taking into account the statistical significance of the
various neutrino signals.
\subsection{ Reactions with angular dependence}
Neutrino-electron scattering
occurs for all flavors of neutrinos and antineutrinos, and is detected
by observing the recoil electrons with kinetic energy $T$ .
The scattering angle is
dictated by the kinematics and is given by
\begin{equation}
\cos\alpha = \frac{E_{\nu} + m_e}{E_{\nu}}
\left( \frac{ T}{T + 2m_e} \right)^{1/2}\,.
\end{equation}
With threshold of about $T_{\rm min} = 5$ MeV, the recoil electrons will
be sharply forward scattered, i.e., pointing away from the supernova,
with the combined
average $\langle \cos \alpha \rangle$ = 0.98, corresponding to about
$11^{\circ}$. However, multiple scattering
will smear the \v{C}erenkov cone,
resulting in a one-sigma width of $\sim 25^\circ$.
In order to evaluate the pointing ability of this signal
we have to take into account
the finite statistics, the two-dimensional form of the resulting distribution,
and the presence of the unavoidable background. The background worsens
the pointing ability from the simple expectation by a factor
\cite{point} $C(R)$
\begin{equation}
\delta x = \frac{\sigma}{\sqrt{N_S}} \times C(R);
~{\rm with} ~ C(R) \approx \sqrt{1 + 4 R} ~,
\end{equation}
where $R$ is the ratio (at the peak)
of the flat background and the signal with $N_S$ events.
For SK and SNO the background reduction factor is $C(R) \simeq 2 - 3$,
and with our standard supernova parameters
we find that the one-sigma error based
on the neutrino-electron scattering will
be about $5^{\circ}$ in SK and about $20^{\circ}$ in SNO.
This is by far the
most accurate pointing ability at our disposal.
The reaction with the most events is $\bar{\nu}_e + p
\rightarrow e^+ + n$, with $\simeq 10^4$ events expected in SK, and
$\simeq 400$ events expected in the light water of SNO.
In \v{C}erenkov detectors one can determine the direction
of the positrons, whose angular distributions
with respect to the direction of the neutrino beam is of the form
\begin{equation}
\frac{d N}{d \cos\alpha} = \frac{N}{2} \left(1 + a \cos\alpha \right)\,.
\label{eq:acos}
\end{equation}
It is relatively easy to show that the error
in the pointing ability for $N$ observed events in this case is given by
\begin{equation}
\delta (\cos\alpha) =
\frac{2}{|a|} \frac{1}{\sqrt{N}}\,.
\end{equation}
Since, in general $a = a(E_\nu)$, we have to investigate further
the neutrino energy dependence of this coefficient, and perform
the necessary energy averaging.
\begin{figure}[h]
\vspace*{13pt}
\begin{center}
\mbox{\epsfig{figure=pos60.eps,width=8cm}}
\vspace*{13pt}
\fcaption{Upper panel: total cross section for $\bar{\nu}_e + p
\rightarrow e^+ + n$; bottom panel: $\langle \cos\theta \rangle$ for
the same reaction; both as a function of the antineutrino energy. The
solid line is the ${\cal O}(1/M)$ result and the short-dashed line is
the ${\cal O}(1)$ result. The long-dashed line is the result of
Eq.(3.18) of Ref.~\protect\cite{LS}, and the dot-dashed line contains our
threshold modifications to the same.}
\label{fig:cos}
\end{center}
\end{figure}
In the limit where the nucleon mass $M$ is taken to be infinite, i.e.,
zeroth order in $1/M$ (${\cal O}(1)$),
the asymmetry coefficient $a$ is independent
of $E_\nu$ and is given simply by the competition of
the non-spin-flip (Fermi) and spin-flip (Gamow-Teller) contributions,
and is
\begin{equation}
a^{(0)} = \frac{f^2 - g^2}{f^2 + 3g^2} \simeq -0.10 ~~;~~
f = 1, g = 1.26 ~,
\label{eq:a0}
\end{equation}
and thus the angular distribution of the positrons is weakly backward.
However, $a(E_\nu)$ is substantially modified when weak magnetism
and recoil corrections of ${\cal O}(1/M)$ are included. It turns out \cite{angul}
that the inclusion to this order gives a very accurate formula for
$\langle \cos \theta \rangle$. This quantity and the total
cross section are shown in Fig. \ref{fig:cos},
evaluated in various approximations (see \cite{angul}).
At high energies, the formula (3.18) of \cite{LS}, valid to all orders
in $1/M$, but neglecting the threshold effects, is applicable. One can see in
Fig. \ref{fig:cos} that the dot-dashed line smoothly interpolates between the
correct low energy and high energy behaviour.
As far as the pointing ability of the
$\bar{\nu}_e + p \rightarrow e^+ + n$ reaction is concerned,
due to the rather small angular asymmetry (and its energy
dependence) we estimate that the uncertainty
$\delta (\cos\alpha) \simeq 0.2$ even for the high statistics
detector like SK. Nevertheless, it would be important and useful
to use this additional information constraining the supernova direction.
\subsection{ Triangulation}
For two detectors separated by a distance $d$, there will be a delay
between the arrival times of the neutrino pulse. The magnitude of the
delay $\Delta t$ depends upon the angle $\theta$ between the supernova
direction and the axis connecting the two detectors. Given a measured
time delay $\Delta t$, the unknown angle $\theta$ and its error
are then:
\begin{equation}
\cos\theta = \frac{\Delta t}{d} ~~;~~
\delta(\cos\theta) = \frac{\delta(\Delta t)}{d}\,.
\end{equation}
Thus two detectors define a cone along their axis with opening
$\cos\theta$ and thickness $2 \times \delta(\cos\theta)$ in which the
supernova can lie. Obviously, in order to have a reasonable pointing
accuracy from triangulation, one will need $\delta(\Delta t) \ll d$.
(The Earth diameter is $d \approx 40$ ms.)
Following \cite{point} I discuss whether an appropriate time delay can be
defined, and what its error would likely be.
Basically, the question can be reduced to the following problem in statistics:
given $N$ events of duration $\tau$, is the uncertainty $\delta(\Delta t)$
equal to $\tau/N$ (i.e., the interval between events) or the much larger
$\tau/\sqrt{N}$?
The answer, of obvious practical significance, requires a degree of subtlety.
Let us model, as before, the time dependence of the neutrino pulse (i.e. the
supernova luminosity $L(t)$) by two exponentials, the increasing sharp rise
with time constant $\tau_1$ and the slow decay with the time constant $\tau_2$
($\tau_1 \ll \tau_2$). Now take the limit (unrealistic) of zero risetime
($\tau_1 \rightarrow 0$). Then, in fact, the first answer is applicable, i.e.,
$\delta(\Delta t) \rightarrow \tau_2/N$. One would then simply
determine the arrival time of the first event in each detector,
and the triangulation would be feasible, though still not very accurate.
But any finite leading edge, or background, would invalidate this picture.
Moreover, we know that the leading edge has a finite duration related
to the shock propagation time in the supernova. The best strategy then
is to try to determine the rather sharp point of maximum rate $t_0$.
The error in its determination depends on the duration of the leading
edge $\tau_1$, which can be measured in the largest detector, and
on the number of events $N_1$ in the leading edge for the given detector,
\begin{equation}
(\delta t_0)_{min} ~\approx ~ \frac{\tau_1}{\sqrt{N_1}} ~.
\end{equation}
At the same time, the number of events $N_1$ in the leading edge
depends somewhat indirectly on the total duration of the pulse,
since $N_1 \approx N \tau_1/\tau_2$. For the existing detectors,
this leads to rather large uncertainty, $\delta(\cos\theta) \approx 0.5$.
Nevertheless, if there will be several large detectors
available in not too distant future, triangulation
would offer another handle to the supernova localization, besides
the obvious benefit of the false alarm elimination by the
coincidence requirement.
\section{Conclusions}
In this talk, which is based on the results of Refs. \cite{SK,SNO,point,angul},
I have shown that:
\begin{itemize}
\item The supernova signal caused by $\nu_{\tau} + \nu_{\mu}$
and their antiparticles can be isolated.
\item By measuring the average arrival time difference of the neutral
and charged current events, one will be able to (conservatively) determine
the upper limit for $m_{\nu_{\tau}}$ of 30-50 eV, representing
an improvement by $10^6$ when compared to the existing limits.
\item Neutrino electron scattering can be used for pointing with
accuracy of about $5^{\circ}$.
\item $\bar{\nu}_e + p \rightarrow e^+ + n$ can be also used
for crude pointing, provided the correct differential cross section
is used. (Remembering that the naive formula suggests that the
positron are slightly backward while in reality they should be
slightly forward.)
\item Triangulation appears to be difficult if the supernova signal
is going to last more than one second. But it would be useful if
more than two detectors (and even better if they are going to be large)
will participate in the warning network.
\end{itemize}
\section{Acknowledgements}
The collaboration with John Beacom is gratefully acknowledged.
This work was supported in part by the US Department of Energy under
Grant No. DE-FG03-88ER-40397.
|
\section{Introduction}
The angular size - redshift relation, $\Theta (z)$, is a
kinematic test which potentially may discriminate the several
cosmological models proposed in the literature. As widely
known,
because of the spacetime curvature, the expanding universe acts
gravitationally
as a lens of large focal length. Though nearby objects are not
affected, a fixed angular size of an extragalactic source is
initially seen decreasing up to a minimal value, say, at a
critical redshift
($z_m$), after which increasing for higher redshifts. The
precise
determination of $z_m$, or equivalently, the corresponding
minimal
angular size value $\Theta(z_m)$, may constitute a powerful
tool in
the search for deciding which are the more realistic world
models. This
lensing effect was first predicted by Hoyle, originally aiming
to distinguish
between the steady-state and Einstein-de Sitter cosmologies \cite{hoyle}.
Later on, the accumulated evidences against the steady state
(mainly from
CMBR) have put it aside, and more recently, the same is
occurring with the
theoretically favoured critical density FRW model \cite{krauss2,Riess,Perlmutter,Alcaniz,Roos}.
The data concerning the angular size - redshift
relation are until nowadays somewhat controversial, specially
because
they envolve at least two kinds of observational dificulties.
First, any
large redshift object may have a wide range of proper sizes,
and, second,
evolutionary and selection effects probably are not negligible.
The $\Theta(z)$ relation for some extended
sources samples seems to be quite imcompatible with the
predictions of the standard FRW model when the latter effects
are not taken into account \cite{Sand,Kapa1,Kapa2}.
There have also been some claims that
the best fit model for the observed distribution of high
redshifts extended
objects is provided by the standard Einstein-de
Sitter universe ($q_o={1 \over 2}$, $\Omega_\Lambda=0$) with
no significant evolution \cite{Bucha}.
Parenthetically, these results are in contradiction with
recent observations
from type Ia supernovae, which seems to ruled out world models
filled only by baryonic matter, and more generally,
any model with positive deceleration parameter \cite{Riess,Perlmutter}. The same happens with
the corresponding bounds using the ages of old high redshift
galaxies \cite{Alcaniz,Dunlop,Krauss3}.
The case for compact radio sources is also of great interest.
These objects
are apparently less sensitive to evolutionary effects since
they are
short-lived ($\sim 10^{3} yr$) and much smaller than their host
galaxy. Initially, the
data from a sample of 82 objects gave remarkable suport for
the Einstein-de Sitter Universe \cite{Kell}. However,
some analysis suggest that Kellerman has not really detected a
significant increasing beyond the minimum \cite{Dabro,Step,Cool}. Some
authors have also argued that models where $\Theta(z)$
diminishes and after a given $z$ remains constant may also
provide a good fit to Kellerman's data. In particular, by
analysing a subset of 59 compact sources within the same
sample,
Dabrowski et al. (1995) found that no useful bounds on the
value of the
deceleration parameter $q_o$ can be derived. Further, even
considering that
Euclidean angular sizes ($\Theta \sim z^{-1}$) are excluded at
99$\%$
confidence level, and that the data are consistent with
$q_o=1/2$, they
apparently do not rule out extreme values of the deceleration
parameter as
$q_{o} \sim 5$ \cite{Step}. More recently, based
in a more
complete sample of data, which include the ones originally
obtained by
Kellermann, it was argued that the $\Theta(z)$ relation may be
consistent with any model of the FRW class with deceleration
parameter $\leq 0.5$ \cite{Gurv}.
In this context, we discuss here how the critical redshift
giving the turn-up in angular sizes is determined for any
expanding cosmology based on the FRW geometry. An analytical
expression quite convenient for numerical evaluation is
derived. The approach is exemplified for three different
models of current cosmological interest: (i) open matter
dominated FRW universe (OCDM), (ii) flat FRW type models with
cosmological
constant ($\Lambda$CDM), (iii) the class of scalar field
cosmologies (SF)
proposed by Ratra and Peebles \cite{Ratra}. Hopefully, the results
derived
here may be useful near future, when more accurate data become
available.
\section{The method}
Let us now consider the FRW line element $(c=1)$
\begin{equation}
ds^2 = dt^2 - R^{2}(t) [d\chi^{2} + S^{2}_{k}(\chi) (d
\theta^2 +
\rm{sin}^{2} \theta d \phi^{2})] \quad ,
\end{equation}
where $\chi$, $\theta$, and $\phi$ are dimensionless comoving
coordinates, $R(t)$ is the scale factor, and $S_{k}(\chi)$
depends on
the curvature parameter ($k=0$, $\pm 1$). The later function is
defined
by one of the following forms: $S_k (\chi) = \rm{sinh} (\chi)$,
$\chi$,
$\rm{sin} \chi$, respectively, for open, flat and closed
Universes.
In this background, the angular size-redshift relation for a
rod of intrinsic length $D$ is easily obtained by integrating
the spatial part of the above expression for $\chi$ and $\phi$
fixed. One finds
\begin{equation}
\theta(z) = {D (1 + z) \over R_{o}S_{k}(\chi)} \quad .
\end{equation}
The dimensionless coordinate $\chi$ is given by
\begin{equation}
\chi(z) = {1 \over H_o R_o} \int_{(1 + z)^{-1}}^{1} {dx \over x
E(x)}
\quad ,
\end{equation}
where $x = {R(t) \over R_o} = (1 + z)^{-1}$ is a convenient
integration variable.
For the three kinds of cosmological models considered here
(OCDM, $\Lambda$CDM and SF) the
dimensionless function $E(x)$ assume one of the following
forms:
\begin{equation}
E_{FRW}(x) = \left[1 - \Omega_{M} + \Omega_{M}
x^{-1}\right]^{{1
\over 2}} \quad ,
\end{equation}
\begin{equation}
E_{\Lambda}(x) = \left[(1 - \Omega_{\Lambda}) x^{-1} +
\Omega_{\Lambda}x^{2}\right]^{{1 \over 2}} \quad ,
\end{equation}
\begin{equation}
E_{SF}(x) = \left[(1 - \Omega_{\phi})x^{-1} +
\Omega_{\phi}x^{{4-\alpha}
\over {2 + \alpha}}\right]^{{1}\over{2}} \quad ,
\end{equation}
where $\Omega_{M} = {{8 \pi G\rho_{M}} \over 3 H_{o}^{2}}$,
$\Omega_{\Lambda} = {\Lambda \over 3H_{o}^{2}}$ and
$\Omega_{\phi} =
{8 \pi G\rho_{\phi} \over 3 H_{o}^{2}}$, are the present day
density
parameters associated with the matter component, cosmological
constant and the scalar field $\phi$, respectively. Notice that
equations (5) and (6) become identical if one takes $\alpha =
0$ in
the later, thereby showing that the scalar field model proposed
by
Ratra and Peebles may kinematically be equivalent to a flat
$\Lambda$CDM cosmology.
The redshift $z_{m}$ at which the angular size takes the
minimum
value is the one cancelling out the derivative of $\Theta$ with
respect to $z$.
Hence, from (2) we have the condition
\begin{equation}
S_k (\chi_m) = (1 + z_m)S'_k (\chi_m) \quad ,
\end{equation}
where $S'_k (\chi) = {\partial S_{k} \over \partial
\chi}{\partial
\chi \over \partial z}$, a prime denotes differentiation with
respect
to $z$ and by definition $\chi_{m}= \chi(z_{m})$. To proceed
further,
observe that (3) can readily be differentiated yielding,
respectively, for the standard FRW (matter dominated),
$\Lambda$CDM
and scalar field cosmologies
\begin{equation}
(1 + z_{m})\chi'_{m} = {(R_o H_o)^{-1} \over \left[1 - \Omega_M
+
\Omega_M (1 + z_m)\right]^{{1 \over 2}}} = (R_o H_o)^{-1}
F(\Omega_{M}, z_m) \quad ,
\end{equation}
\begin{equation}
(1 + z_{m})\chi'_{m} = {(R_o H_o)^{-1} \over \left[(1 -
\Omega_{\Lambda})(1 + z_m) + \Omega_{\Lambda}(1 +
z_m)^{-2}\right]^{{1 \over 2}}} = (R_o H_o)^{-1}
L(\Omega_{\Lambda},
z_m) \quad ,
\end{equation}
\begin{equation}
(1 + z_{m})\chi'_{m} = {(R_o H_o)^{-1} \over \left[(1 -
\Omega_{\phi})(1 + z_m) + \Omega_{\phi}(1 + z_m)^{{\alpha - 4
\over
\alpha + 2}}\right]^{{1 \over 2}}} = (R_o H_o)^{-1}
S(\Omega_{\phi}, \alpha, z_m) \quad .
\end{equation}
Now, inserting the above equations into (7) we find
for the cases above considered
\begin{equation}
{1 \over (1 - \Omega_{M})^{1 \over 2}}{\rm{tanh}}\left[(1 -
\Omega_{M})^{1 \over 2}\int_{(1 + z_m)^{-1}}^{1} {dx \over x
E_{FRW}(x)}\right] =
F(\Omega_{M}, z_m) \quad ,
\end{equation}
\begin{equation}
\int_{(1 + z_m)^{-1}}^{1} {dx \over x E_{\Lambda}(x)} =
L(\Omega_{\Lambda}, z_m) \quad ,
\end{equation}
\begin{equation}
\int_{(1 + z_m)^{-1}}^{1} {dx \over x E_{SF}(x)} =
S(\Omega_{\phi},
\alpha, z_m) \quad .
\end{equation}
The meaning of equations (11)-(13) is self evident. Each one
represents an integro-algebraic equation for the critical
redshift
$z_m$ as a function of the physically meaningful parameters of
the
models. In general, these equations cannot be solved in closed
analytical form for $z_m$. However, as one may check, if we
take the
limit $\Omega_M \rightarrow 1$ in (11), the value $z_m = {5
\over 4}$
is readily achieved, which corresponds to the well known
standard
result for the dust filled FRW flat universe. The interesting
point is
that expressions (11)-(13) are quite convenient for numerical
evaluations. As a matter of fact,
their solutions can straightforwardly be obtained, for
instance, by
programming the integrations using simple numerical recipes in
FORTRAN.
In Fig.~1 we show the diagrams of $z_m$ as a function of the
density parameter for each kind of model. As expected, in the
standard FRW model, the critical redshift starts at $z_m=1.25$
when $\Omega_M$ goes to unity. This value is pushed to the
right direction, that is, it is
displaced to higher redshifts as the $\Omega_M$ parameter is
decreased. For instance, for $\Omega_M=0.5$ and $\Omega_M=0.2$,
we find $z_m=1.58$ and $z_m=2.20$, respectively. In the
limiting case, $\Omega_M \rightarrow 0$, there is
no minimum at all since $z_{m} \to \infty$. This means that the
angular size decreases monotonically as a function of the
redshift.
For the scalar field case, one needs to fix the value of
$\alpha$ in
order to have a bidimensional plot. Given a value of
$\Omega_{\phi}$,
the minimum is also displaced for higher redshifts when the $\alpha$
parameter diminishes. Conversely, for a fixed value of
$\alpha$, the
minimum moves for lower redshifts when $\Omega_\phi$ is
decreased. The limiting case ($\alpha=0$) is fully equivalent
to a
$\Lambda$CDM model. As happens in the limiting case $\Omega_M
\rightarrow 0$ ($\Omega_{\Lambda} = 0$), the minimal value for
$\Theta(z)$
disappears when the cosmological constant contributes all the
energy density
of the Universe, that is, $z_m \rightarrow \infty$ if
$\Omega_M \rightarrow 0$
and $\Omega_{\Lambda} \rightarrow 1$ (in this connection see
also \cite{krauss1}). For the class of models considered in this
paper, the
redshifts having the minimal angular size are displayed for
several values of $\Omega_M$ and $\alpha$ in Table 1. As can be
seen there, the critical
redshift at which the angular size is a minimal cannot alone
discriminate
between world models since different scenarios may provide the
same $z_m$
value. However, when combinated with other tests, some
interesting
constraints on the cosmological models can be obtained.
For example, when $\Omega_\phi$ is bigger than $0.55$, the model proposed by Ratra and Peebles yields a $z_m$
between the standard FRW flat model and the $\Lambda$CDM cosmology. Then,
suposing that the universe is really accelerating today ($q_{o} < 0$), as
indicated recently by measurements using type Ia supernovae \cite{Riess,Perlmutter}, and by considering
the results by Gurvits et al. \cite{Gurv}, i.e., that the data are
compatible with
$q_{o} \leq 0.5$, the Ratra and Peebles models with $0 < \alpha
\leq 4$ seems
to be more in accordance with the angular size data for compact
radio sources
than the $\Lambda$CDM model.
\begin{table}[t]
\begin{center}
\begin{tabular}{rrlll}
\hline
\multicolumn{1}{c}{$\Omega_{m}$ ($z_{m}$)}&
\multicolumn{1}{c}{$\Omega_{\Lambda}$ ($z_{m}$)}&
\multicolumn{1}{c}{$\Omega_{\phi} (\alpha = 2)$ ($z_{m}$)}&
\multicolumn{1}{c}{$\Omega_{\phi} (\alpha = 4)$ ($z_{m}$)}&
\multicolumn{1}{c}{$\Omega_{\phi} (\alpha = 6)$ ($z_{m}$)}\\
\\
\hline
1.0 (1.25)& 1.0 ($\infty$)& 1.0 (2.16)& 1.0 (1.72)& 1.0
(1.57)\\
\\
0.8 (1.35)& 0.8 (1.76)& 0.8 (1.65)& 0.8 (1.53)& 0.8 (1.46)\\
\\
0.7 (1.41)& 0.7 (1.60)& 0.7 (1.55)& 0.7 (1.47)& 0.7 (1.42)\\
\\
0.5 (1.58)& 0.5 (1.44)& 0.5 (1.42)& 0.5 (1.38)& 0.5 (1.36)\\
\\
0.2 (2.20)& 0.2 (1.31)& 0.2 (1.30)& 0.2 (1.30)& 0.2 (1.29)\\
\hline
\end{tabular}
\caption{Critical redshift $Z_m$ in OCDM, $\Lambda$CDM, and scalar field cosmologies for some selected values of the density parameters.}
\end{center}
\end{table}
\pagebreak
\begin{figure}
\vspace{.2in}
\centerline{\psfig{figure=angsn.ps,width=3truein,height=3truein}
\hskip 0.1in}
\caption{Critical redshift $Z_m$ as a function of the density
parameter in open, ${\Lambda}\rm{CDM}$ and scalar field cosmologies. Solid
curve is the prediction for a model with nonnull cosmological constant. The same curve is also obtained as a limiting case ($\alpha \rightarrow 0$) of the scalar field cosmology proposed by Ratra and Peebles}
\end{figure}
It is worth notice that the same procedure may be applied when
evolutionary
and/or selection effects due to a linear size-redshift or to a
linear size-luminosity dependence are taken into account. As
widely believed, a plausible way of standing for such effects
is to consider that the intrinsic linear size has a similar
dependence on the redshift as the coordinate dependence, i.e.,
$D = D_o (1 + z)^{c}$, being $c < 0$ (see, for instance,
\cite{Bucha} and Refs. therein). In this case,
equations (11)-(13) are still valid but the functions
$F(\Omega_{M}, z_m)$, $L(\Omega_{\Lambda}, z_m)$, and
$S(\Omega_{\phi}, \alpha, z_m)$ must be divided by a factor $(1
+ c)$. The displacement of $z_{m}$ relative to the case with no evolution
($c = 0$) due to the effects cited above may be unexpectedly
large. For example, if one takes $c = -0.8$ as found by
Buchalter et al. \cite{Bucha}, the redshift of the minimum angular
size for the Einstein-de Sitter case ($\Omega_{M} = 1$) moves
from $z_{m} = 1.25$ to $z_{m} = 11.25$. In particular, this
explains why the data of Gurvits et al. \cite{Gurv}, although
apparently in agreement with the Einstein-de Sitter universe,
do not show clear evidence for a minimal angular size close to
$z = 1.25$, as should be expected for this model.
\hspace{0.3cm}
{\bf Acknowledgments:}This work was partially
supported by the project Pronex/FINEP (No. 41.96.0908.00) and
Conselho Nacional de Desenvolvimento Cient\'{\i}fico e
Tecnol\'ogico - CNPq (Brazilian Research Agency).
|
\section{Introduction: Wide-angle scattering in QCD}
Although attempts to apply perturbative QCD to wide-angle elastic hadron
scatterings have been undertaken in a number of papers [1-10], explicit
predictions have been available only for elastic processes involving external
photons, such as $\gamma + \gamma\rightarrow$ hadrons, Compton scattering
of hadrons, etc.
Predictions based on perturbative QCD rest on three premises: 1)
hadronic interactions become weak at small invariant separation
$r\ll \Lambda_{QCD}^{-1}$; 2) the perturbative expansion in
$\alpha_s(Q)$ is well-defined; and 3) factorization, implying that all
effects of collinear singularities, confinement, non-perturbative
interactions and bound state dynamics can be isolated at large momentum
transfer in terms of the (process independent) structure functions
$G_{i/H}(x,Q),$ fragmentation functions $D_{H/i}(z,Q)$ or, in the case of
exclusive processes, distribution amplitudes $\Phi_H(s_i,Q).$
Consequently the hadronic scattering amplitude takes the form
\begin{equation}
A=\int\prod_H\phi_H(x_i,Q) T(x_i,p_H;Q)[dx_i]~,
\label{eq1}
\end{equation}
where $\Phi(x_i,Q)$ is a universal distribution amplitude which gives
the probability amplitude for finding the valence $q\bar q$ or $qqq$
in the hadronic wave function collinear up to the scale
$Q=\sqrt{s\over 2}$, and $T$ is the hard scattering amplitude for valence
quark collisions.
The technical complication which has made particularly difficult to
compute the behavior of hadron-hadron amplitudes is the possibility of
multiple scatterings. The standard factorized form for the elastic
scattering of hadrons $\{i\}$ is
\begin{equation}
A_1(s,t)=\int^1_0\prod_{i=1}^4[dx_i]\phi_i(x_i)T(\{x_i\},s,t)~,
\label{eq2}
\end{equation}
where $x_i$ represents collectively the fractional momenta of hadron $i$
carried by its valence partons.
According to this concept, all of the partons collide in a small region of
the space-time of typical dimension $Q^{-1}$. The relevant contribution to
the amplitude behaves according to the dimensional counting \cite{BF73,
mtm73}, i.e.
\begin{equation}
A_1(s,t)\sim({\mu^2\over s})^{n/2-2}f_1(s/t)
\label{eq3}
\end{equation}
for $n$ partons participating in the hard scattering, $\mu$ representing
hadronic mass scales, which make the amplitude dimensionless.
An extension of this "single-scattering" scenario is the (double) "independent-
scattering" picture, due to P. Landshoff \cite{Lan74}, in which two
pairs of partons
scatter independently off two scattering centers. According to this
picture, the lowest order diagrams contribute with
\begin{equation}
A_m(s,t)\sim\biggl({\mu^2\over s}\biggr)^{(n-m+1)/2-2}f_m(s/t)~,
\label{eq4}
\end{equation}
where $m$ is the number of independent scatterings.
If so, the multiple scattering should dominate in the case of wide angle
scattering.
A solution to this problem was pointed out in Refs.~\cite{LB80} and
\cite{LP81}, where it was shown that the Sudakov logarithms associated with
the rescattering diagrams do not cancel. In the leading logarithmic
approximation they exponentiate to suppress the typical double scattering
contribution by a factor
\begin{displaymath}
exp(-const\ln Q^2\ln(\ln Q^2))~,
\end{displaymath}
characteristic of the Sudakov suppression in QCD.
More quantitatively \cite{M81},
\begin{equation}
A_2\sim{1\over{Q^4}}\biggl({Q\over \Lambda_{QCD}}\biggr)^{1-2c\ln (1/r)}\,,
\label{eq5}
\end{equation}
where
\begin{displaymath}
r=2c/(1+2c)~,~~~\ \ \ \ c=32/(33-2n_f)~,
\end{displaymath}
and $n_f$ is the number of flavors. Interestingly, for $n_f=3$ the
power turns out to be $Q^{-3.8},$ nearly the same as the dimensional
counting power $Q^{-4}$ in the single-scattering scenario.
Higher order diagrams were calculated e.g. in Ref.~\cite{FSZ89}, however soon
it became evident that even the first order QCD correction involves
an immense number of Feynman diagrams, so further attempts to go
beyond the simple quark counting rule were abandoned.
It may be that perturbative QCD is not the relevant (or not
the only physically interesting) expansion of the wide-angle
scattering amplitude. Recent developments in M-branes (see e.g.
Ref.~cite{polchinski}) may open new prospects in the realization of a
hypothetical duality between small and large distances (or,
equivalently, large- and small-angle scattering). The search for a
relevant expansion parameter is of crucial importance on this way.
In this paper we are solving an "inverse problem": we use the known
explicit expression of the dual amplitude with Mandelstam
analyticity (DAMA), that has correct wide angle scaling
behavior. By identifying it with that resulting from the quark
counting rules, we then calculate two sub-leading terms in the
expansion of the known full dual amplitude and study the behavior
of the resulting series.
\section{ Wide angle behavior of the dual amplitude with
Mandelstam analyticity}
\label{sec1}
Wide-angle scaling behavior within the $S-$matrix approach
was discussed in Ref.~\cite{coon78}, where
by means of a logarithmic Regge trajectory an interpolation from
the ``soft`` Regge behavior to the ``hard`` scaling regime was
suggested. The motivation of the logarithmic trajectory came
from earlier papers \cite{alik75}, where
a class of dual models requiring a logarithmic trajectory was suggested.
The logarithmic asymptotic behavior of the trajectory and the large angle
scaling behavior are uniquely related also in a different class
of dual models, called dual amplitudes with Mandelstam analytics
(DAMA) \cite{jenk79,bcj80}. The link between this class of models in the
scaling
limit and the parton model in the infinite momentum frame was
studied in Ref.~\cite{sch73}. In all those papers only the leading
asymptotic
($s,|t|\rightarrow \infty,\ s/t = const$) term was treated. The results of
different approaches vary in such details as the form of the
scaling violation (normally, logarithmic), the form of the angular
dependence $f(\theta )$ and the way active quarks are counted.
In this paper we calculate the sub-leading terms in the pre-asymptotic
(larger $s$ and $|t|$) behavior of DAMA. Since the model is realistic
enough in the sense that it satisfies the general requirements
of the theory (see Refs.~\cite{jenk79,bcj80}), we believe that our result is
universal and thus it may be used as a guide e. g. in QCD calculations.
Apart from the leading term, we have explicitly calculated two
more sub-leading terms. Our technique allows further calculations
of still higher orders, but the obtained first three terms of
the series already show a regular trend that may be interpreted
as the expansion in the running coupling constant
$g(s) \sim 1/\ln s,$ valid at large
$s$ and $|t|$. This situation takes place for anyone trajectory
with the logarithmic asymptotic.
The aim of the present paper is two-fold. First, by identifying the
leading term of the asymptotic (wide-angle) expansion of DAMA with that
derived from perturbative QCD \cite{LB80} we tentatively assume that the
DAMA in the wide angle asymptotic region is equivalent to the
asymptotically free regime in QCD. With this identification in mind, we
calculate within DAMA corrections to the leading term in the hope that their
form may give some insight into the relevant corrections in perturbative
QCD that are known to be very complicated.
Clearly, the above identity has the chance to be true only in the vicinity
of the wide angle region (small distances), where perturbative calculations
are assumed to be still valid.
The second aspect is purely phenomenological. Since, however, the experimental
situation in the wide-angle region did not change for almost two decades,
we are left with the earlier fits to the data.
Let us now calculate the ``perturbative`` expansion of DAMA.
We write the elastic scattering amplitude for spinless particles
in the following symmetric form \cite{bcj80}:
\begin{equation}
A(s,t,u)=C(s-u)\lbrack D(s,t)-D(u,t)\rbrack~,
\label{eq6}
\end{equation}
where $C$ is a constant and
\begin{equation}
D(s,t)=\int\limits_{0}^{1}dx\left( \frac{x}{g} \right) ^{-\alpha
(s' )} \left( \frac{1-x}{g} \right) ^{-\alpha (t' )}~.
\label{eq7}
\end{equation}
Here
$s' =s(1-x),\ \ \ t' =tx$
and $g$ is a dimensionless parameter, $g>1$. Only one, leading trajectory
was included and it was chosen in a simple, but
representative form:
\begin{equation}
\alpha (s)=\alpha _{0} -\gamma \ln \left( \frac{1+\beta \sqrt{s_{0} -s}}
{1+\beta \sqrt{s_{0} } } \right)~,
\label{eq8}
\end{equation}
that account both for the threshold and the asymptotic behavior and
is nearly linear for very small $|s|,|s|<<s_{0}\,.$
For simplicity we have included only the leading trajectories in both
channels: the Pomeron trajectory in the $t$-channel and the exotic
trajectory in the $s$-channel. While the parameters of the Pomeron
trajectory are well known, only a little is known about the exotic
trajectory. Fortunately, this has no substantial effect on our results, since
our goal is the functional form of the series and its individual
terms rather than fits to the data. Given the scarcity of the data
and the freedom available in the model, the wide-angle behavior of
DAMA cannot be determined completely.
Let us consider the asymptotic behavior of Eq.~(\ref{eq7}) in the limit
$s,\left| t\right| \rightarrow \infty ,\ s/t=const.$
For the Regge trajectories we have
\begin{equation}
\alpha(s)=\alpha (0)-\frac{\gamma }{2} \ln \delta ^{2}+ i\pi \frac{\gamma}
{2}- \frac{\gamma }{2}\ln s=-a-\lambda~,
\label{eq9}
\end{equation}
\begin{equation}
\alpha (t)=\alpha (0)-\frac{\gamma }{2} \ln \delta ^{2}-
\frac{\gamma } {2}\ln\left(-\frac{t}{s}\right)-\frac{\gamma }{2}\ln s=
-b-\lambda~,
\label{eq10}
\end{equation}
with
\begin{displaymath}
a=-\alpha (0)+\frac{\gamma }{2} \ln \delta ^{2} -i\pi
\frac{\gamma }{2}~,~~~~
b=-\alpha (0)+\frac{\gamma }{2} \ln \delta ^{2} +\frac{\gamma }{2}
\ln(-\frac{t}{s})~,
\end{displaymath}
\begin{equation}
\delta=\frac{\beta \sqrt{s_{0}}}{1+\beta \sqrt{s_{0} } }~,~~~~~~
\lambda = \frac{\gamma }{2} \ln s~.
\label{eq11}
\end{equation}
From here on, $s, t, u$ will be dimensionless variables, measured
in units of $s_0$.
In this domain the saddle point method can be used to calculate
the integral in Eq.~(\ref{eq7}) \cite{mag97}. To do this we can rewrite
Eq.~(\ref{eq7}) in the following form
\begin{equation}
D(s,t)=(2g)^{-a-b-2\lambda} g^{\gamma \ln 2} {1\over 2}
\int\limits_{-1}^{1}g(u)~e^{\lambda f(u)} du~,
\label{eq12}
\end{equation}
where we have changed the variable $x$ to $u$, $x=(1-u)/2$, and introduced
new functions:
\begin{equation}
g(u)=(1-u)^{\tilde{a}}(1+u)^{\tilde{b}}e^{\gamma \ln{\frac{1-u}{2}}
\ln{\frac{1+u}{2}}}~,
\label{eq13}
\end{equation}
\begin{equation}
f(u)=\ln (1-u^2) ~,
\label{eq14}
\end{equation}
\begin{equation}
\tilde{a}=a-\frac{\gamma}{2}\ln g ~,\ \ \ \ \ \tilde{b}=b-\frac{\gamma}{2}\ln g~.
\label{eq15}
\end{equation}
We see now that $f(u)$ has a sharp maximum at the saddle point $u_0=0$.
We quote the explicit expression for the saddle point expansion in the
\ref{ap_coef}.
Using formulas from this Appendix
we obtain the power series for $D(s,t)$ in \ref{powerD}. It reads
\begin{equation}
D(s,t)\approx \frac{A_1 s^{-\gamma \ln 2g}}{\sqrt{\gamma \ln s}}
\left(-\frac{t}{s}\right)^{-\frac{\gamma }{2} \ln 2g}\left\{1+
\frac{h_1(\tilde{a},\tilde{b})}{\gamma \ln s}+\frac{h_2(\tilde{a},\tilde{b})}{(\gamma \ln s)^2}\right\}~,
\label{eq16}
\end{equation}
where $A_1,\ h_1, \ h_2$ are given by the expressions
(\ref{b5}, \ref{b8}, \ref{b9}).
The expression for $D(u,t)$ can be calculated in a similar way (see
Eq.~(\ref{b10}) in \ref{powerD}).
In the kinematical region
$s, \left| t\right| \rightarrow \infty,\ t/s=const$
we can use the substitutions
\begin{equation}
t\approx -s\cdot \sin ^{2} \left( \theta/2 \right)~,~~~~~~
u\approx -s\cdot \cos ^{2} \left( \theta/2 \right)~.
\label{eq17}
\end{equation}
Substituting the results for $D(s,t)$ and $D(u,t)$ into Eq.~(\ref{eq6}) and
changing the variables we get the expression
for the full amplitude as a function of the $s$ and $\theta$ variables
(see Eq.~(\ref{b14}) in \ref{powerD}):
\begin{equation}
A(s,\theta)\approx \frac{C A s^{-N}}{\sqrt{\gamma \ln s}} f(\theta)
I(s,\theta )~,
\label{eq18}
\end{equation}
where $A,\ N,\ f(\theta),\ I(s,\theta )$ are given by the expressions
(\ref{b13}, \ref{b15} - \ref{b17}).
To summarize, we have expanded the wide-angle scattering amplitude in a
power series of $1/\ln{s}$ and have evaluated explicitly the
coefficients of the first two terms (beyond the leading one).
\section{Comparison with the data and discussion of the results}
\label{sec2}
New experimental data on wide-angle scatterings are not likely to
appear any more because of the simple reason that as energy increases more
particles tend to fly in the forward direction and there is no chance
to detect e.g. the proton-proton differential cross section at
$90^{\circ}$ for, say, $\sqrt s> 10\ GeV$. ``Wide angles'', of course,
extend beyond $90^{\circ}$. Still the complication due to the huge number
of Born diagrams contributing to large angle exclusive reactions \cite{LB80},
overwhelming the contribution due to the Landshoff pinch singularity \cite{Lan74},
will remain for long topical in this field. We use the data given in the
compilation of \cite{LP73} to fix the scale. The errors, quoted in the
original papers (see Ref.~\cite{LP73} and references therein), are
typically about 10
overall normalization factor, the "quark counting power" in the cross
section being set equal to $N=4$ in the case of proton-proton
cross section, in agreement with the data \cite{LP73, LB80}
(see Fig.~\ref{fig1}).
\begin{figure}[htb]
\centerline{\psfig{figure={pryamye_0.eps},height=11.0cm}}
\caption{\footnotesize Cross section $\frac{d\sigma }{dt}$ for {\bf pp}
$\rightarrow$ {\bf pp} scattering
at various center of mass scattering angles. Both axes are in logarithmic
scale. Stars denote the experimental points from Ref.~\cite{LP73}.
The straight lines correspond to a falloff of
$\sim 1/s^{10}$. They are calculated according to the power series for the
scattering amplitude, discussed above ($\frac {d\sigma}{dt}=
\frac{4\pi}{(s s_0)^2}|A(s,\theta)|^2$), with the following set of
parameters:
$\alpha_0 = 1$,
$N=4,\ \gamma = 2.84 \ (g=2.9),$ $\beta=0.05 \ GeV^{-1}$, $C=2.7\cdot10^{-14}
GeV^{-2}$ and $s_0 = 4m_\pi^2$.
}
\label{fig1
\end{figure}
Our main goal is the behavior of the scaling-violating corrections
to the leading term obeying quark counting rules. Fig. \ref{fig2} shows the
relative contribution of these terms.
We draw the correction power series:
\begin{displaymath}
J(s,\theta)=|I(s,\theta)|^2 \approx 1+ 2\frac{Re(f_1(\theta)/Z(\theta))}
{\gamma \ln s}+
2\frac{Re(f_2(\theta)/Z(\theta))}{(\gamma \ln s)^2}+\frac{|f_1(\theta)
/Z(\theta)|^2}{(\gamma \ln s)^2}
\end{displaymath}
\begin{equation}
+O({1\over \lambda^{3}})~,
\label{eq19}
\end{equation}
where $f_1(\theta),~f_2(\theta),~Z(\theta)$ are given by expressions
(\ref{b18} - \ref{b20}).
We can see that the corrections
are quite large for small $s$, especially for angles close to $90^0$. That is
not a surprise, since the lowest order of our
expansion is valid for large $s$ ($\gamma \ln s /2 >> 1$). In the experimental
energy interval the corrections give factor $4-6$ to the cross sections and
should not be neglected. This was missed in the references \cite{jenk79, bcj80}.
Moreover we find that the
corrections are very sensitive to variations of $\beta$ and $\gamma$.
\begin{figure}[htb]
\centerline{\psfig{figure={popravki_0.eps},height=11.0cm}}
\caption{\footnotesize
The corrections $J(s,\theta)$, given by
Eq.~(\ref{eq19}), to the differential cross section $\frac{d\sigma }{dt}$
for {\bf pp} $\rightarrow$ {\bf pp} scattering .
We have used the same values of parameters as in Fig.~\ref{fig1}:
$\alpha_0 = 1$, $N=4,\ \gamma = 2.84\ (g=2.9),$ $\beta=0.05 \ GeV^{-1}$
and $s_0 = 4m_\pi^2$, coming from the comparison with the data.}
\label{fig2}
\end{figure}
\vskip 1cm
{\bf Acknowledgment}
V.K. Magas is thankful for the hospitality extended to him by the Bogolyubov
Institute for Theoretical Physics in Kiev where part of this work was done.
|
\section*{Introduction}
Every representation $\alpha:G\to\Aut B({\mathcal H})$ of a discrete group $G$
as automorphisms of the algebra $B({\mathcal H})$
of bounded linear operators on a complex Hilbert space ${\mathcal H}$
determines a unique element of $H^2(G,\field{T})$:
each $\alpha_s$ is of the form $\Ad U_s$ for some unitary operator $U_s$,
and $U_sU_t = \omega(s,t)U_{st}$ determines a $2$\nobreakdash- cocycle
$\omega$ whose cohomology class is independent of the choice
of implementing unitaries.
If instead one starts with an action
of a semigroup $P$ as endomorphisms of $B({\mathcal H})$,
then, as described in \cite[Remark~2.3]{fowrae},
the cohomological obstruction one encounters is a
{\em product system\/} over $P$:
roughly speaking, a collection
$E = \{E_s:s\in P\}$ of complex Hilbert spaces together with
an associative multiplication which implements unitary isomorphisms
$E_s\otimes E_t \to E_{st}$.
Product systems were first defined by Arveson \cite{arv}
in his study of one-parameter semigroups
of endomorphisms of $B({\mathcal H})$.
In this note we continue the investigations of
\cite{dinhjfa, dinhjot, laca, fowrae, fowler}
into the $C^*$\nobreakdash- algebras
associated with discrete product systems.
Each of the algebras studied in these previous papers
can be regarded as a Toeplitz algebra:
unit vectors of the product system
correspond to isometries in the universal $C^*$\nobreakdash- algebra,
and finite orthonormal sets of vectors from a fiber $E_s$
map to isometries whose range projections are orthogonal
and whose sum is strictly less than the identity.
Here we consider only product systems $E$ whose fibers
are finite-dimensional,
and in our universal $C^*$\nobreakdash- algebra ${\mathcal O}_E$
the range projections of the isometries associated
with an orthonormal basis for any fiber sum to the identity.
Hence ${\mathcal O}_E$ is generated by a collection
of Cuntz algebras, one for each element of the underlying semigroup $P$;
the relations between these Cuntz algebras
are determined by the multiplication in $E$.
We think of ${\mathcal O}_E$ as a semigroup version
of the twisted group algebra $C^*(G,\omega)$.
We begin in Section~\ref{section:OE} by establishing
the existence of ${\mathcal O}_E$ and discussing a few examples.
In particular one can recover as ${\mathcal O}_E$ the Cuntz algebras
(Remark~\ref{remark:cuntz algebra}) and the irrational rotation
algebras (Remark~\ref{remark:irrational}).
In Section~\ref{section:abelian} we develop some basic results
regarding the structure of ${\mathcal O}_E$
when the underlying semigroup $P$ embeds in an abelian group;
this is a key hypothesis in most of our results.
In Section~\ref{section:simplicity} we prove our main result,
Theorem~\ref{theorem:main}:
if no two fibers of $E$ have the same dimension,
then ${\mathcal O}_E$ is simple and purely infinite.
For the lexicographic product systems of \cite{fowrae}
our Theorem~\ref{theorem:lexicographic} gives a sharper simplicity result:
${\mathcal O}_E$ is simple if and only if the dimension function is injective.
This shows in particular that ${\mathcal O}_E$ need not be simple
even if the only one-dimensional fiber of $E$ is the fiber over
the identity of $P$.
For product systems over $\field{N}^2$ we show in Theorem~\ref{theorem:N2}
that ${\mathcal O}_E$ is either
simple or isomorphic to ${\mathcal O}_l\otimes C(\field{T})$, where $l$ is determined
by the dimensions of the fibers over $(1,0)$ and $(0,1)$.
Finally, in Theorem~\ref{theorem:nuclearity} we generalize a technique
of Laca \cite{laca} to show that ${\mathcal O}_E$ is nuclear
if $E$ admits a twisted unit.
\section{The Cuntz algebra of a product system}\label{section:OE}
Suppose $P$ is a countable discrete semigroup with identity $e$.
A {\em product system\/} over $P$ is a family $p:E\to P$
of nontrivial separable complex Hilbert spaces $E_t:= p^{-1}(t)$
which is endowed with an associative multiplication $E\times E\to E$
in such a way that $p$ is a semigroup homomorphism,
and such that for every $s,t\in P$ the map
$x\otimes y\in E_s\otimes E_t \mapsto xy\in E_{st}$
extends to a unitary isomorphism.
We also insist that $\dim E_e = 1$,
so that $E$ has an identity $1\in E_e$ \cite[Lemma~1.3]{fowrae}.
A {\em representation\/} of $E$ in a $C^*$\nobreakdash- algebra $B$
is a map $\phi:E\to B$ which satisfies
\begin{itemize}
\item[(i)] $\phi(xy) = \phi(x)\phi(y)$ for every $x,y\in E$; and
\item[(ii)] $\phi(y)^*\phi(x) = (x\mid y) \phi(1)$
if $p(x) = p(y)$.
\end{itemize}
This definition is slightly different than that found in
\cite{fowrae}, where representations were on Hilbert space
and condition (ii) was that
$\phi(y)^*\phi(x) = (x\mid y) 1$ if $p(x) = p(y)$.
The main advantage to our definition is that it allows us to
consider the trivial map $E\to\{0\}$ as a representation,
without regarding $\{0\}$ as a unital $C^*$\nobreakdash- algebra.
If $\phi \ne 0$, then $\phi(1)$ is a nontrivial projection
which serves as an identity for the $C^*$\nobreakdash- subalgebra
$C^*(\phi)\subseteq B$ generated by $\phi(E)$;
moreover, $\phi$ restricts to
an isometric linear map on each of the fibers of $E$
\cite[p.8]{arv}.
For nontrivial $\phi$,
condition (ii) implies that each $\phi(x)$ is a multiple of an isometry
in $C^*(\phi)$;
indeed, if ${\mathcal B}_t$ is an orthonormal basis for $E_t$,
then $\{\phi(f): f\in {\mathcal B}_t\}$ is a family of isometries with
mutually orthogonal range projections.
In this note we will consider only product systems whose fibers
are finite-dimensional, and we are primarily interested in representations
$\phi:E\to B$ which satisfy
\begin{equation}\label{eq:cuntz rep}
\sum_{f\in {\mathcal B}_t} \phi(f)\phi(f)^* = \phi(1)
\qquad\text{for all $t\in P$.}
\end{equation}
We call a representation $\phi$ which satisfies \eqref{eq:cuntz rep}
a {\em Cuntz representation\/}.
While nontrivial representations of $E$ always exist
\cite[Lemma~1.10]{fowrae},
it is not clear that every product system admits a nontrivial
Cuntz representation.
\begin{prop}\label{prop:OE}
Let $E$ be a product system over $P$ of finite-dimensional Hilbert spaces.
There is a pair $({\mathcal O}_E,i_E)$ consisting of a $C^*$\nobreakdash- algebra ${\mathcal O}_E$
and a Cuntz representation $i_E:E\to{\mathcal O}_E$ with the following properties:
\textup{(a)} for every Cuntz representation $\phi$ of $E$,
there is a homomorphism $\phi_*$ of ${\mathcal O}_E$,
called the {\em integrated form of $\phi$\/},
such that $\phi_*\circ i_E = \phi$; and
\textup{(b)} ${\mathcal O}_E$ is generated as a $C^*$\nobreakdash- algebra by $i_E(E)$.
\noindent The pair $({\mathcal O}_E,i_E)$ is unique up to canonical isomorphism.
\end{prop}
\begin{remark}\label{remark:trivial}
We emphasize that ${\mathcal O}_E$ is trivial if $E$ does not admit
a nontrivial Cuntz representation.
In all other cases ${\mathcal O}_E$ is unital and $i_E$ is isometric.
\end{remark}
\begin{proof}[Proof of Proposition~\ref{prop:OE}]
The Proposition can be proved by modifying the standard argument
of, for example, \cite[Proposition~1.3]{fowrae2}.
Briefly, let $S$ be a collection of Cuntz representations on
Hilbert space which are cyclic (i.e. generate a $C^*$\nobreakdash- algebra
which admits a cyclic vector), and such that every cyclic
Cuntz representation is unitarily equivalent to an element of $S$.
Define $i_E := \bigoplus_{\phi\in S} \phi$ and ${\mathcal O}_E := C^*(i_E)$.
Condition (a) holds because every Cuntz representation of $E$
on a Hilbert space decomposes as the direct sum of a zero representation
and a collection of cyclic representations,
and the uniqueness assertion follows from a standard argument.
\end{proof}
\begin{example}\label{example:lexicographic}
{\em (Lexicographic product systems.)\/}
For any product system $E$ whose fibers are finite-dimensional,
$s \mapsto \dim E_s$ is a semigroup homomorphism from $P$
to the multiplicative positive integers $\field{N}^*$.
We call this homomorphism the {\em dimension function\/} of $E$.
Let $d:P\to\field{N}^*$ be an arbitrary homomorphism.
In \cite[Examples~1.4(E2)]{fowrae} it was shown how one can construct
a product system $E(d)$, called the {\em lexicographic product system
determined by $d$\/}, whose dimension function is $d$:
for each $n$ let $\{\delta_0,\dots,\delta_{n-1}\}$
be the canonical basis for $\field{C}^n$,
take
\[
E(d) := \bigsqcup_{s\in P} \{s\}\times \field{C}^{d(s)},
\]
and define multiplication on basis vectors using
the lexicographic order on
$\{0,\dots, d(r) - 1\} \times \{0, \dots, d(s) - 1\}$; that is,
\[
(r,\delta_j)(s,\delta_k) := (rs,\delta_{jd(s) + k})
\qquad\text{for $0 \le j \le d(r) - 1$ and $0 \le k \le d(s) - 1$.}
\]
Multiplication in $E(d)$ is defined by extending bilinearly.
There is a distinguished Cuntz representation of $E(d)$
on $L^2(\field{T})$. Suppose $r\in P$ and $0 \le j \le d(r) - 1$.
Let $S(r,\delta_j)$ be the operator on $L^2(\field{T})$
whose value on a vector $\xi\in L^2(\field{T})$ is given by
\[
S(r,\delta_j)\xi(e^{2\pi it})
:= \begin{cases}
d(r)^{1/2} \xi(e^{2\pi i(td(r) - j)})
& \text{if $t\in \bigl[ \frac j{d(r)}, \frac{j+1}{d(r)} \bigr)$} \\
0 & \text{otherwise.} \end{cases}
\]
It is easy to see that $S(r,\delta_j)$ is an isometry,
and that
\[
\sum_{j=0}^{d(r) - 1} S(r,\delta_j)S(r,\delta_j)^* = 1 = S(1).
\]
Since
\begin{align*}
& S(r,\delta_j)S(s,\delta_k)\xi(e^{2\pi it}) \\
& \qquad = \begin{cases}
d(r)^{1/2} S(s,\delta_k)\xi(e^{2\pi i(td(r) - j)})
& \text{if $t\in \bigl[ \frac j{d(r)}, \frac{j+1}{d(r)} \bigr)$} \\
0 & \text{otherwise} \end{cases} \\
& \qquad = \begin{cases}
d(r)^{1/2} d(s)^{1/2} \xi(e^{2\pi i((td(r) - j)d(s) - k)})
& \text{if $td(r) - j\in \bigl[ \frac k{d(s)}, \frac{k+1}{d(s)} \bigr)$} \\
0 & \text{otherwise} \end{cases} \\
& \qquad = \begin{cases}
d(rs)^{1/2} \xi(e^{2\pi i(td(rs) - (jd(s) + k))})
& \text{if $t \in \bigl[
\frac {jd(s) + k}{d(rs)}, \frac{jd(s) + k + 1}{d(rs)} \bigr)$} \\
0 & \text{otherwise} \end{cases} \\
& \qquad = S(rs,\delta_{jd(s)+k})\xi(e^{2\pi it}),
\end{align*}
we have
$S(r,\delta_j)S(s,\delta_k) = S(rs,\delta_{jd(s)+k})
= S((r,\delta_j)(s,\delta_k))$.
Hence defining
\[
S(r,x) := \sum_{j=0}^{d(r)-1} (x \mid \delta_j) S(r,\delta_j)
\qquad\text{for $(r,x)\in E(d)$}
\]
gives a Cuntz representation $S:E(d)\to B(L^2(\field{T}))$.
\end{example}
\begin{remark}\label{remark:cuntz algebra}
Fix $n \ge 2$ and let $d$ be the homomorphism $a\in\field{N} \mapsto n^a\in\field{N}^*$.
Then ${\mathcal O}_{E(d)}$ is the Cuntz algebra ${\mathcal O}_n$.
\end{remark}
\begin{example}
{\em (One-dimensional product systems.)\/}
Suppose $\omega$ is a multiplier of $P$;
that is, a function $\omega:P\times P\to\field{T}$
such that $\omega(e,e) = 1$ and
\[
\omega(r,s)\omega(rs,t) = \omega(r,st)\omega(s,t)
\qquad\text{for all $r,s,t\in P$.}
\]
Then
$(r,z)(s,w) := (rs,\omega(r,s)zw)$
for $r,s\in P$ and $z,w\in\field{C}$
defines an associative multiplication which gives
$P\times\field{C}$ the structure of a product system;
we write $(P\times\field{C})^\omega$ for this product system.
It is easy to see that every product system whose fibers
are one-dimensional is isomorphic to $(P\times\field{C})^\omega$
for some multiplier $\omega$.
Suppose $P$ is an Ore semigroup; that is,
suppose $P$ is cancellative and satisfies $Pr\cap Ps \ne\emptyset$
for every pair $r,s\in P$. (For example, every cancellative commutative
semigroup has this property.)
By \cite[Theorem~1.1.2]{laca2}, $P$ can be embedded
in a group $G$ with $P^{-1}P = G$,
and by \cite[Theorem~1.2.2]{laca2} there is a multiplier
$\omega'$ of $G$ which extends $\omega$.
Define $l^\omega:(P\times\field{C})^\omega\to B(\ell^2(G))$
by
\[
l^\omega(r,z)\xi(s) := z\omega'(r,r^{-1}s)\xi(r^{-1}s)
\qquad\text{for $\xi\in\ell^2(G)$ and $s\in G$.}
\]
It is routine to check that each $l^\omega(r,1)$ is unitary
and that $l^\omega$ is multiplicative.
Hence $l^\omega$ is a Cuntz representation of $(P\times\field{C})^\omega$.
\end{example}
\begin{remark}\label{remark:irrational}
Let $\theta\in(0,1)$ be irrational,
let $\omega:\field{N}^2\times\field{N}^2\to\field{T}$ be the multiplier
$\omega((a,b),(c,d)) := e^{2\pi i\theta bc}$,
and let $E = (\field{N}^2\times\field{C})^\omega$.
Then $U := i_E((0,1),1)$ and $V := i_E((1,0),1)$
are unitaries which generate ${\mathcal O}_E$ and satisfy
$UV = e^{2\pi i\theta} VU$,
so ${\mathcal O}_E$ is the irrational rotatation algebra ${\mathcal A}_\theta$.
\end{remark}
\begin{example}
{\em (Twisting.)\/} If $E$ is a product system over $P$
and $\omega$ is a multiplier of $P$,
then $(x,y)\in E\times E \mapsto \omega(p(x),p(y))xy$
defines a multiplication on $E$
which also gives $E$ the structure of a product system;
we write $E^\omega$ for this new system,
and say $E$ has been {\em twisted by $\omega$\/}.
If $\phi$ is a Cuntz representation of $E$
on a Hilbert space ${\mathcal H}$, then
\[
\phi^\omega(x) := \phi(x)\otimes l^\omega(p(x),1)
\qquad\text{for $x\in E^\omega$}
\]
defines a Cuntz representation $\phi^\omega$
of $E^\omega$ on ${\mathcal H}\otimes\ell^2(G)$.
\end{example}
The results obtained in the previous examples allow us to state:
\begin{prop}\label{prop:nontrivial}
Every twisted lexicographic product system over an Ore semigroup
admits a nontrivial Cuntz representation.
\end{prop}
\section{Product systems over abelian semigroups}
\label{section:abelian}
The following Proposition collects some results concerning the
structure of ${\mathcal O}_E$ when $P$ embeds in an abelian group.
\begin{prop}\label{prop:abelian}
Suppose $P$ is a subsemigroup of a countable abelian group $G$
and $E$ is a product system over $P$ of finite-dimensional Hilbert spaces.
\textup{(1)}
Let $\phi$ be a Cuntz representation of $E$,
let $s,t\in P$, and let ${\mathcal B}_s$ and ${\mathcal B}_t$
be orthonormal bases for $E_s$ and $E_t$, respectively.
Then for any $x'\in E_s$ and $y'\in E_t$ we have
\[
\phi(y')^*\phi(x') = \sum_{x\in {\mathcal B}_s} \sum_{y\in {\mathcal B}_t} (x'y \mid y'x) \phi(x)\phi(y)^*.
\]
\textup{(2)} ${\mathcal O}_E = \clsp\{i_E(x)i_E(y)^*: x,y\in E\}$.
\textup{(3)} There is a strongly continuous action $\gamma:\widehat G\to\Aut{\mathcal O}_E$,
called the {\em gauge action\/},
such that
\[
\gamma_\lambda(i_E(x)) = \lambda(p(x))i_E(x)
\qquad\text{for all $\lambda\in\widehat G$ and $x\in E$.}
\]
\textup{(4)} Let $m$ be Haar measure on $\widehat G$.
Then
\[
\Phi(b) := \int_{\widehat G} \gamma_\lambda(b)\,dm(\lambda)
\qquad\text{for $b\in{\mathcal O}_E$}
\]
defines a faithful conditional expectation $\Phi$
of ${\mathcal O}_E$ onto ${\mathcal O}_E^\gamma$,
the fixed-point algebra of the gauge action.
\textup{(5)} ${\mathcal O}_E^\gamma
= \clsp\{i_E(x)i_E(y)^*: x,y\in E,\ p(x) = p(y)\}$.
\end{prop}
\begin{proof}
(1) We use \eqref{eq:cuntz rep} to calculate
\begin{align*}
\phi(y')^*\phi(x')
& = \Bigl( \sum_{x\in {\mathcal B}_s} \phi(x)\phi(x)^* \Bigr)
\phi(y')^*\phi(x')
\Bigl( \sum_{y\in {\mathcal B}_t} \phi(y)\phi(y)^* \Bigr) \\
& = \sum_{x\in {\mathcal B}_s} \sum_{y\in {\mathcal B}_t}
\phi(x)\phi(y'x)^*\phi(x'y)\phi(y)^* \\
& = \sum_{x\in {\mathcal B}_s} \sum_{y\in {\mathcal B}_t}
(x'y \mid y'x)\phi(x)\phi(y)^*,
\end{align*}
noting that $p(y'x) = p(x'y)$ since $P$ is abelian.
(2) Take $\phi = i_E$ in (1) to see that
the linear span of monomials of the form $i_E(x)i_E(y)^*$
is closed under multiplication, and hence a ${}^*$\nobreakdash- algebra.
(3) For each $\lambda\in\widehat G$
the map $x\in E \mapsto \lambda(p(x))i_E(x)$
is a Cuntz representation, and hence integrates to
an endomorphism $\gamma_\lambda$ of ${\mathcal O}_E$.
Since $\gamma_\lambda \circ \gamma_{\lambda^{-1}}$
and $\gamma_{\lambda^{-1}} \circ \gamma_\lambda$
are both the identity on ${\mathcal O}_E$, $\gamma_\lambda$
is an automorphism. Obviously $\gamma$ is a group
homomorphism, and its continuity follows from a straightforward
$\epsilon/3$ argument.
(4) This is a standard result about automorphic actions of
discrete abelian groups.
(5) Since $\Phi(i_E(x)i_E(y)^*) = \delta_{p(x),p(y)}i_E(x)i_E(y)^*$,
(5) follows from (2) and the continuity of $\Phi$.
\end{proof}
Our next goal is to give an abstract characterization of the
fixed-point algebra ${\mathcal O}_E^\gamma$. First some notation.
If $s,t\in P$, $S\in{\mathcal K}(E_s)$ and $T\in{\mathcal K}(E_t)$,
write $S\otimes T$ for the operator on $E_{st}$ which satisfies
\[
S\otimes T(xy) = (Sx)(Ty)
\qquad\text{for $x\in E_s$ and $y\in E_t$.}
\]
Write $1^t$ for the identity operator on $E_t$.
Define a relation $\preceq$ on $P$ by $s \preceq t$ if and only if $t\in sP$,
and observe that $\preceq$ is a preorder on $P$;
that is, it is reflexive and transitive.
Since $P$ is commutative, for every pair $s,t\in P$
we have $st \in sP\cap tP$, and hence $(P,\preceq)$ is upwardly directed.
Now $S\mapsto S\otimes 1^t$
is a unital embedding of ${\mathcal K}(E_s)$ in ${\mathcal K}(E_{st})$,
and since multiplication in $E$ is associative we have
$S\otimes 1^{tr} = (S\otimes 1^t)\otimes 1^r$.
Hence $S\in{\mathcal K}(E_s) \mapsto S\otimes 1^t\in{\mathcal K}(E_{st})$
is a directed system of $C^*$\nobreakdash- algebras,
and we can define
\[
{\mathcal F}_E := \varinjlim {\mathcal K}(E_s).
\]
Since each ${\mathcal K}(E_s)$ is simple, so is ${\mathcal F}_E$.
(It is not hard to see that ${\mathcal F}_E$ is a UHF algebra.)
Let $\iota_s$ be the canonical embedding of ${\mathcal K}(E_s)$ in ${\mathcal F}_E$.
\begin{prop}\label{prop:FE}
There is a unique homomorphism $i_{\mathcal F}:{\mathcal F}_E\to{\mathcal O}_E$
which satisfies
\[
i_{\mathcal F}(\iota_{p(x)}(\rankone xy)) = i_E(x)i_E(y)^*
\qquad\text{whenever $p(x) = p(y)$,}
\]
where $\rankone xy$ is the rank-one operator
$z\in E_{p(x)} \mapsto (z\mid y)x$.
The image of ${\mathcal F}_E$ is precisely ${\mathcal O}_E^\gamma$.
If $E$ admits a nontrivial Cuntz representation,
then $i_{\mathcal F}$ is injective.
\end{prop}
\begin{proof}
For each $s\in P$, there is a unique homomorphism $\sigma_s:{\mathcal K}(E_s)\to{\mathcal O}_E$
such that $\sigma_s(\rankone xy) = i_E(x)i_E(y)^*$ for all $x,y\in E_s$.
Let ${\mathcal B}_t$ be an orthonormal basis for $E_t$. Since
\begin{align*}
\sigma_{st}((\rankone xy)\otimes 1^t)
& = \sigma_{st}\Bigl( \sum_{f\in {\mathcal B}_t} (\rankone xy) \otimes (\rankone ff) \Bigr)
= \sigma_{st}\Bigl( \sum_{f\in {\mathcal B}_t} (\rankone {xf}{yf} \Bigr) \\
& = \sum_{f\in {\mathcal B}_t} i_E(xf)i_E(yf)^*
= i_E(x)\Bigl( \sum_{f\in {\mathcal B}_t} i_E(f)i_E(f)^* \Bigr) i_E(y)^* \\
& = i_E(x)i_E(y)^* = \sigma_s(\rankone xy),
\end{align*}
we deduce that $\sigma_{st}(S\otimes 1^t) = \sigma_s(S)$ for all $S\in{\mathcal K}(E_s)$,
and hence there is a homomorphism $i_{\mathcal F}:{\mathcal F}_E\to{\mathcal O}_E$
such that $i_{\mathcal F}\circ\iota_s = \sigma_s$ for every $s\in P$.
From Proposition~\ref{prop:abelian}(5) it is obvious that
$i_{\mathcal F}$ maps ${\mathcal F}_E$ onto ${\mathcal O}_E^\gamma$.
If $E$ admits a nontrivial Cuntz representation then $i_E$ is nonzero
(see Remark~\ref{remark:trivial}), hence each $\sigma_s$
is nonzero, and finally $i_{\mathcal F}$ is nonzero.
Since ${\mathcal F}_E$ is simple, we deduce that $i_{\mathcal F}$ is injective.
\end{proof}
\section{Simplicity and pure infiniteness}
\label{section:simplicity}
\begin{theorem}\label{theorem:main}
Let $G$ be a countable abelian group,
let $P$ be a subsemigroup of $G$ which contains the identity,
and let $E$ be a product system over $P$ of finite-dimensional Hilbert spaces
which admits a nontrivial Cuntz representation.
If the dimension function $s \mapsto \dim E_s$ is injective,
then ${\mathcal O}_E$ is simple and purely infinite.
\end{theorem}
To prove the Theorem we require a technical lemma.
For the Proposition which precedes it, see
\cite[Proposition~2.7]{arv} and \cite[Proposition~1.11 and Lemma~3.6]{fowrae}.
\begin{prop}\label{prop:alpha}
Let $E$ be a product system over $P$ of finite-dimensional Hilbert spaces,
and let $\phi$ be a Cuntz representation of $E$
in a unital $C^*$\nobreakdash- algebra $B$ such that $\phi(1) = 1$.
For each $s\in P$ there is a unital endomorphism $\alpha^\phi_s$ of $B$
which satisfies
\[
\alpha^\phi_s(b) = \sum_{f\in {\mathcal B}_s} \phi(f)b\phi(f)^*
\qquad\text{for all $b\in B$}
\]
whenever ${\mathcal B}_s$ is an orthonormal basis for $E_s$.
Moreover, $\alpha_e$ is the identity, and
\[
\alpha^\phi_{st}(b)\phi(x) = \phi(x)\alpha^\phi_t(b)
\qquad\text{for all $x\in E_s$.}
\]
\end{prop}
\begin{lemma}\label{lemma:kill}
Let $E$ be a product system satisfying the hypotheses
of Theorem~\ref{theorem:main},
and let $\phi$ be a Cuntz representation of $E$
in a unital $C^*$\nobreakdash- algebra $B$ such that $\phi(1) = 1$.
Suppose $x_1$, \dots, $x_n$,
$y_1$, \dots, $y_n\in E$ satisfy $p(x_i) \ne p(y_i)$ for $1 \le i \le n$.
Let $c\in P$ be such that $p(x_i)^{-1}c \in P$ and $p(y_i)^{-1}c\in P$
for every $i$.
Then there is a unit vector $w\in E$ such that
\begin{equation}\label{eq:w}
\alpha^\phi_c(\phi(w)\phi(w)^*)\phi(x_i)\phi(y_i)^*
\alpha^\phi_c(\phi(w)\phi(w)^*) = 0
\qquad\text{for $1 \le i \le n$.}
\end{equation}
\end{lemma}
\begin{proof}
Define $s_i := p(x_i)^{-1}c$, $t_i := p(y_i)^{-1}c$,
and for each $r\in P$ let ${\mathcal B}_r$ be an orthonormal basis for $E_r$.
Suppose $Q\in \phi_*({\mathcal O}_E)$;
we will eventually take $Q = \phi(w)\phi(w)^*$.
By Proposition~\ref{prop:alpha} we have
\begin{multline*}
\alpha^\phi_c(Q)\phi(x_i)\phi(y_i)^*\alpha^\phi_c(Q)
= \phi(x_i)\alpha^\phi_{s_i}(Q)\alpha^\phi_{t_i}(Q)\phi(y_i)^* \\
= \sum_{f'\in {\mathcal B}_{s_i}} \sum_{g'\in {\mathcal B}_{t_i}}
\phi(x_i)\phi(f')Q\phi(f')^*\phi(g')Q\phi(g')^*\phi(y_i)^*,
\end{multline*}
and by Proposition~\ref{prop:abelian}(1)
\[
Q\phi(f')^*\phi(g')Q
= \sum_{g\in {\mathcal B}_{t_i}} \sum_{f\in {\mathcal B}_{s_i}}
(g'f \mid f'g) Q\phi(g)\phi(f)^*Q,
\]
so it suffices to find a unit vector $w\in E$ such that
\begin{equation}\label{eq:sufficient}
\phi(w)^*\phi(g)\phi(f)^*\phi(w) = 0
\qquad\text{whenever $(f,g)\in {\mathcal B}_{s_i}\times {\mathcal B}_{t_i}$
for some $i$.}
\end{equation}
Let $(f_l, g_l)_{l=1}^m$ be an enumeration of these pairs.
We claim that for each $j\in \{0, \dots, m\}$
there is a unit vector $v_j \in E$ such that
\begin{equation}\label{eq:v}
\phi(v_j)^*\phi(g_l)\phi(f_l)^*\phi(v_j) = 0
\qquad\text{for $1 \le l \le j$.}
\end{equation}
Given the claim, $w := v_m$ satisfies \eqref{eq:sufficient},
and hence \eqref{eq:w}, completing the proof.
The claim is vacuous when $j = 0$: taking any unit vector $v_0\in E$ works.
Suppose inductively that
there exists $k \le m-1$ such that \eqref{eq:v} holds when $j = k$.
Let $r := p(v_k)$, $s := p(f_{k+1})$, and $t := p(g_{k+1})$.
Since $s \ne t$, by hypothesis $\dim E_s \ne \dim E_t$,
and we can assume without loss of generality that $\dim E_s < \dim E_t$.
We have
\begin{align*}
& \phi(v_k)^*\phi(g_{k+1})\phi(f_{k+1})^*\phi(v_k) \\
& \quad = \sum_{f\in {\mathcal B}_s} \sum_{g\in {\mathcal B}_t} \sum_{v\in {\mathcal B}_r}
\phi(g)\phi(g)^*\phi(v_k)^*\phi(g_{k+1})
\phi(v)\phi(v)^*\phi(f_{k+1})^*\phi(v_k)
\phi(f)\phi(f)^* \\
& \quad = \sum_{f\in {\mathcal B}_s} \sum_{g\in {\mathcal B}_t} \sum_{v\in {\mathcal B}_r}
(g_{k+1}v \mid v_kg)(v_kf \mid f_{k+1}v) \phi(g)\phi(f)^* \\
& \quad = \sum_{f\in {\mathcal B}_s} \phi(u_f)\phi(f)^*,
\end{align*}
where
\[
u_f := \sum_{g\in {\mathcal B}_t} \sum_{v\in {\mathcal B}_r}
(g_{k+1}v \mid v_kg)(v_kf \mid f_{k+1}v) g \in E_t.
\]
Since $\dim E_s < \dim E_t$, there is a unit vector
$v'\in E_t$ which is orthogonal to each $u_f$,
and taking $v_{k+1} := v_kv'$ gives \eqref{eq:v} for $j = k+1$.
\end{proof}
\begin{proof}[Proof of Theorem~\ref{theorem:main}]
Suppose $B$ is a $C^*$\nobreakdash- algebra and
$\pi:{\mathcal O}_E\to B$ is a nonzero homomorphism.
We will show that $\pi$ is injective,
thus establishing the simplicity of ${\mathcal O}_E$.
It does not harm to assume that $\pi$ is surjective,
so that $B$ is unital and $\pi(1) = 1$.
Recall from Proposition~\ref{prop:abelian}(4) that there
is a faithful expectation $\Phi$ of ${\mathcal O}_E$ onto ${\mathcal O}_E^\gamma$,
given by averaging over the orbits of the gauge action.
We will show that there is an expectation $\Phi_\pi$
of $\pi({\mathcal O}_E)$ onto $\pi({\mathcal O}_E^\gamma)$ such that
\begin{equation}\label{eq:spatial Phi}
\Phi_\pi\circ\pi = \pi\circ\Phi.
\end{equation}
To see that this implies that $\pi$ is injective,
suppose $b\in\ker\pi$.
Then
$\pi\circ\Phi(b^*b) = \Phi_\pi\circ\pi(b^*b) = 0$,
so $\Phi(b^*b)\in {\mathcal O}_E^\gamma\cap\ker\pi$.
But ${\mathcal O}_E^\gamma$ is simple and contains the identity of ${\mathcal O}_E$,
so ${\mathcal O}_E^\gamma\cap\ker\pi = \{0\}$.
Thus $\Phi(b^*b) = 0$, and we deduce from the faithfulness
of $\Phi$ that $b = 0$.
By Proposition~\ref{prop:abelian}(2),
finite sums of the form $b = \sum i_E(x_i)i_E(y_i)^*$
are dense in ${\mathcal O}_E$.
Hence to prove the existence of an expectation $\Phi_\pi$
satisfying \eqref{eq:spatial Phi}, it suffices to fix such an
element $b$ and show that
$\lVert \pi(b) \rVert \ge \lVert \pi(\Phi(b)) \rVert$.
Thus with $\phi := \pi\circ i_E$, we must show that
\begin{equation}\label{eq:inequality}
\lVert \sum \phi(x_i)\phi(y_i)^* \rVert
\ge
\Bigl\lVert \sum_{p(x_i) = p(y_i)} \phi(x_i)\phi(y_i)^* \Bigr\rVert.
\end{equation}
Let $c := \prod_i p(x_i)p(y_i)$.
Since $\phi$ is a Cuntz representation of $E$,
the relations \eqref{eq:cuntz rep}
allow us to assume that $p(x_i) = p(y_i) = c$
whenever $p(x_i) = p(y_i)$.
Since $\phi(1) = 1$,
Lemma~\ref{lemma:kill} applies
and provides a unit vector $w\in E$ such that
\[
\alpha^\phi_c(\phi(w)\phi(w)^*)
\phi(x_i)\phi(y_i)
\alpha^\phi_c(\phi(w)\phi(w)^*)
= 0
\qquad\text{whenever $p(x_i) \ne p(y_i)$.}
\]
Let $Q := \phi(w)\phi(w)^*$.
Then
\begin{equation}\label{eq:Q inequality}
\begin{split}
\lVert \sum \phi(x_i)\phi(y_i)^* \rVert
& \ge \lVert \alpha_c(Q) \sum \phi(x_i)\phi(y_i)^* \alpha_c(Q) \rVert \\
& = \Bigl\lVert \alpha_c(Q)\sum_{p(x_i) = p(y_i)}
\phi(x_i)\phi(y_i)^*\alpha_c(Q)
\Bigr\rVert \\
& = \Bigl\lVert \sum_{p(x_i) = p(y_i)} \phi(x_i)Q\phi(y_i)^*
\Bigr\rVert.
\end{split}
\end{equation}
Since $\rankone xy \mapsto \phi(x)Q\phi(y)^*$
and $\rankone xy \mapsto \phi(x)\phi(y)^*$
both extend linearly to nontrivial homomorphisms
of the simple algebra ${\mathcal K}(E_c)$,
we have
\[
\Bigl\lVert \sum_{p(x_i) = p(y_i)} \phi(x_i)Q\phi(y_i)^* \Bigr\rVert
= \Bigl\lVert \sum_{p(x_i) = p(y_i)} \phi(x_i)\phi(y_i)^* \Bigr\rVert.
\]
Combining this with \eqref{eq:Q inequality}
gives \eqref{eq:inequality},
completing the proof of simplicity.
Our proof that ${\mathcal O}_E$ is purely infinite is an adaptation of
the proof of \cite[Proposition~5.3]{bprs}.
Let $A$ be a hereditary subalgebra of ${\mathcal O}_E$; we will show that
$A$ has an infinite projection.
Fix a positive element $a\in A$, scaled so that $\lVert\Phi(a)\rVert = 1$.
Choose a finite sum
$b = \sum i_E(x_i)i_E(y_i)^* \in{\mathcal O}_E$
such that $b \ge 0$ and $\lVert a - b \rVert < 1/4$.
Then $b_0 := \Phi(b)$ is also positive and satisfies $\lVert b_0 \rVert \ge 3/4$.
By applying the relations \eqref{eq:cuntz rep}
we can assume that there exists $c\in P$
such that $p(x_i) = p(y_i) = c$ whenever $p(x_i) = p(y_i)$, and such that
$p(x_i)^{-1}c \in P$ and $p(y_i)^{-1}c \in P$ for all $i$.
Then $b_0$ is a positive element of the algebra
\[
{\mathcal F}_c := \lsp\{i_E(x)i_E(y)^*: p(x) = p(y) = c\},
\]
and hence its image under the canonical isomorphism ${\mathcal F}_c \cong {\mathcal K}(E_c)$
has a unit eigenvector $f\in E_c$ with eigenvalue $\lVert b_0 \rVert$.
It follows that $b_0i_E(f) = \lVert b_0 \rVert i_E(f)$,
and hence the projection $r := i_E(f)i_E(f)^*$ satisfies
$rb_0r = \lVert b_0 \rVert r$.
By Lemma~\ref{lemma:kill}, there is a unit vector $w\in E$ such that
\[
\alpha^{i_E}_c(i_E(w)i_E(w)^*)
i_E(x_i)i_E(y_i)
\alpha^{i_E}_c(i_E(w)i_E(w)^*)
= 0
\]
whenever $p(x_i) \ne p(y_i)$.
Let $q := i_E(fw)i_E(fw)^*$,
noting that $q \le r$ and $q \le \alpha^{i_E}_c(i_E(w)i_E(w)^*)$.
Then
\[
qbq = qb_0q = qrb_0rq
= \lVert b_0 \rVert qrq
= \lVert b_0 \rVert q
\ge \textstyle\frac 34 q,
\]
and since $\lVert a - b \rVert < 1/4$ we thus have
$qaq \ge qbq - \frac 14q \ge \frac 12q$.
It follows that $qaq$ is invertible in $q{\mathcal O}_Eq$. Let $c$ be its inverse,
and define $v := c^{1/2}qa^{1/2}$.
Then $v$ is a partial isometry since $vv^* = c^{1/2}qaqc^{1/2} = q$,
and its initial projection belongs to $A$ since
$v^*v = a^{1/2}qcqa^{1/2} \le \lVert c\rVert a$
and $A$ is hereditary.
Since $v^*v \sim q = i_E(fw)i_E(fw)^* \sim i_E(fw)^*i_E(fw) = 1$
and ${\mathcal O}_E$ contains nonunitary isometries,
we deduce that $v^*v$ is an infinite projection in $A$.
\end{proof}
\section{Simplicity for lexicographic product systems}
For the lexicographic product systems of
Example~\ref{example:lexicographic},
we have the following partial converse to Theorem~\ref{theorem:main}.
\begin{theorem}\label{theorem:lexicographic}
Let $G$ be a countable abelian group,
let $P$ be a subsemigroup of $G$ which contains the identity,
let $d:P\to\field{N}^*$ be a semigroup homomorphism,
and let $E$ be the lexicographic product system determined by $d$.
Then ${\mathcal O}_E$ is simple if and only if $d$ is injective.
\end{theorem}
\begin{proof}
By Proposition~\ref{prop:nontrivial},
$E$ admits a nontrivial Cuntz representation,
so one direction follows from Theorem~\ref{theorem:main}.
For the converse, suppose $d$ is not injective.
Fix $s,t\in P$ such that $s\ne t$ and $d(s) = d(t)$.
We claim that $b := i_E(s,\delta_0) - i_E(t,\delta_0)$
generates a proper ideal ${\mathcal I}\triangleleft{\mathcal O}_E$,
so that ${\mathcal O}_E$ is not simple.
Let $S:E\to B(L^2(\field{T}))$ be the distinguished Cuntz representation
defined in Example~\ref{example:lexicographic}.
Since $S_*(b) = S(s,\delta_0) - S(t,\delta_0) = 0$
we have $b\in\ker S_*$,
and hence ${\mathcal I}$ is not all of ${\mathcal O}_E$.
To see that ${\mathcal I}$ is nonzero,
choose $\lambda\in\widehat G$ such that
$\lambda(s) \ne \lambda(t)$,
and define $T(r,x) := \lambda(r)S(r,x)$ for all $(r,x)\in E$.
Then $T$ is a Cuntz representation of $E$ such that
\[
T_*(b) = T(s,\delta_0) - T(t,\delta_0) = (\lambda(s) - \lambda(t))S(s,\delta_0) \ne 0,
\]
and we deduce that $b$, and hence ${\mathcal I}$, is nonzero.
\end{proof}
When $P = \field{N}^2$ we can say a bit more.
For every $m,n\in\field{N}^*$,
write $E(m,n)$ for lexicographic product system over $\field{N}^2$
determined by the homomorphism $(a,b)\in\field{N}^2 \mapsto m^an^b \in\field{N}^*$.
Note that $E(m,n) \cong E(n,m)$.
\begin{theorem}\label{theorem:N2}
If $m = 1$, then ${\mathcal O}_{E(m,n)} \cong {\mathcal O}_n\otimes C(\field{T})$.
If $m,n \ge 2$ and $\log_m n$ is irrational, then ${\mathcal O}_{E(m,n)}$
is simple.
If $m,n \ge 2$ and $\log_m n$ is rational, then
there exists a unique positive integer $l$
such that $m = l^a$ and $n = l^b$
for some relatively prime positive integers $a$ and $b$;
for this $l$ we have ${\mathcal O}_{E(m,n)} \cong {\mathcal O}_l\otimes C(\field{T})$.
\end{theorem}
The proof of this Theorem is preceded by two Propositions.
The first deals with constructing representations of
lexicographic product systems over $\field{N}^k$,
allowing for $k = \infty$ by defining
$\field{N}^\infty := \bigoplus_{i=1}^\infty \field{N}$.
Let $\{e_a: 1 \le a \le k\}$ be the canonical basis for $\field{N}^k$.
\begin{prop}\label{prop:construct reps}
Let $d:\field{N}^k\to\field{N}^*$ be a semigroup homomorphism,
and let $B$ be a unital $C^*$\nobreakdash- algebra.
Suppose $\{U_{a,i}: 1 \le a \le k,\ 0 \le i \le d(e_a) - 1 \}$
is a set of isometries in $B$ which satisfies
\[
U_{a,i}^*U_{a,j} = 0
\qquad\text{whenever $i \ne j$, and}
\]
\begin{equation}\label{eq:commutation}
U_{a,i}U_{b,j} = U_{b,p}U_{a,q}
\qquad\text{whenever $id(e_b) + j = pd(e_a) + q$.}
\end{equation}
Then there is a unique representation $\phi:E(d) \to B$
such that $\phi(a,\delta_i) = U_{a,i}$,
and $\phi$ is a Cuntz representation if
\begin{equation}\label{eq:cuntz}
\sum_{i=0}^{d(e_a) - 1} U_{a,i}U_{a,i}^* = 1
\qquad\text{for every $a$.}
\end{equation}
\end{prop}
\begin{proof}
Fix $s\in\field{N}^k$ and $m\in\{0,\dots, d(s) - 1\}$.
Our first goal is to define $\phi(s,\delta_m)$.
In any expression of $s$ as an ordered sum
$e_{a_1} + \dotsb + e_{a_l}$ of basis elements
we have $l = \lVert s \rVert_1$.
For each of these finitely-many ordered $l$\nobreakdash- tuples
$(a_1, \dots, a_l)$,
it is easy to see that
there is a unique way of factoring $(s,\delta_m)$
as a product $(e_{a_1}, \delta_{i_1})\dotsb (e_{a_l}, \delta_{i_l})$.
These factorizations can be obtained from one another
by using the commutation relations
\[
(e_a,\delta_i)(e_b, \delta_j) = (e_b, \delta_p)(e_a, \delta_q)
\qquad\text{whenever $id(e_b) + j = pd(e_a) + q$}
\]
on adjacent factors a finite number of times,
and hence \eqref{eq:commutation} implies that the corresponding product
$U_{a_1,i_1}\dotsb U_{a_l,i_l}$
is independent of the factorization;
we define $\phi(s,\delta_m)$ to be this common element of $B$.
It is obvious that $\phi(s,\delta_m)\phi(t,\delta_n) = \phi((s,\delta_m)(t,\delta_n))$,
and defining
\[
\phi(s,x) := \sum_{i=0}^{d(s) - 1} (x \mid \delta_i) \phi(s, \delta_i)
\qquad\text{for $(s,x)\in E(d)$}
\]
gives the desired representation $\phi$.
If \eqref{eq:cuntz} holds, then
\[
\sum_{m=0}^{d(s) - 1} \phi(s,\delta_m)\phi(s,\delta_m)^*
= \sum_{i_1 = 0}^{d(e_{a_1}) - 1} \dotsb
\sum_{i_l = 0}^{d(e_{a_l}) - 1}
U_{a_1,i_1}\dotsb U_{a_l,i_l}U_{a_l,i_l}^*\dotsb U_{a_1,i_1}^*
= 1,
\]
so $\phi$ is a Cuntz representation.
\end{proof}
\begin{prop}\label{prop:factor}
${\mathcal O}_{E(m,n)} \cong {\mathcal O}_{E(m, mn)}$.
\end{prop}
\begin{proof}
Let $E = E(m,n)$ and $F = E(m,mn)$.
Since $\dim F_{(a,b)} = m^a(mn)^b = m^{a+b}n^b = \dim E_{(a+b,b)}$,
we can define $\psi:F\to{\mathcal O}_E$ by
\[
\psi((a,b),x) = i_E((a+b,b),x)
\qquad\text{for all $((a,b),x)\in F$.}
\]
It is easy to see that $\psi$ is a Cuntz representation of $F$,
so there is a homomorphism $\psi_*:{\mathcal O}_F \to {\mathcal O}_E$
such that $\psi_*\circ i_F = \psi$.
Note that for any $(a,b)\in\field{N}^2$ and $j\in\{0,\dots, a^mb^n-1\}$
we have
\begin{align*}
i_E((a,b),\delta_j)
& = i_E((b,0),\delta_i)^*i_E((a+b,b),\delta_{im^an^b + j}) \\
& = \psi((b,0),\delta_i)^*\psi((a,b),\delta_{im^an^b + j});
\end{align*}
since elements of the form $i_E((a,b),\delta_j)$
generate ${\mathcal O}_E$, $\psi_*$ is surjective.
To see that $\psi_*$ is injective, we construct its inverse.
First define
\[
V_{1,i} := i_F((1,0),\delta_i) \in {\mathcal O}_F
\qquad\text{for $0 \le i \le m-1$, and}
\]
\[
V_{2,k} := i_F((0,1),\delta_k) \in {\mathcal O}_F
\qquad\text{for $0 \le k \le mn-1$.}
\]
Note that the $V_{1,i}$'s and $V_{2,k}$'s
are isometries which generate ${\mathcal O}_F$,
that $\sum V_{1,i}V_{1,i}^* = 1 = \sum V_{2,k}V_{2,k}^*$,
and that
\[
V_{1,i}V_{2,k} = V_{2,k'}V_{1,i'}
\qquad\text{whenever $imn + k = k'm + i'$.}
\]
Now define
$U_{1,i} := V_{1,i}$ for $0 \le i \le m-1$ and
\[
U_{2,j} := \sum_{l=0}^{m-1} V_{2,jm+l}V_{1,l}^*
\qquad\text{for $0 \le j \le n-1$.}
\]
It routine to check that the $U_{2,j}$'s are isometries whose range projections
sum to the identity.
Suppose $in + j = pm + q$, where
$0 \le i,q \le m-1$ and $0 \le j,p \le n-1$. Then
\begin{align*}
U_{1,i}U_{2,j}
& = \sum_{l=0}^{m-1} V_{1,i}V_{2,jm + l}V_{1,l}^*
= \sum_{l=0}^{m-1} V_{2,in + j}V_{1,l}V_{1,l}^* \\
& = V_{2,in + m} = V_{2,pm + q}
= \sum_{l=0}^{m-1} V_{2,pm + l}V_{1,l}^*V_{1,q}
= U_{2,p}U_{1,q},
\end{align*}
so by Proposition~\ref{prop:construct reps} there is a
Cuntz representation $\phi:E\to{\mathcal O}_F$ such that
$\phi((1,0),\delta_i) = U_{1,i}$ and $\phi((0,1),\delta_j) = U_{2,j}$.
We check that $\phi_*\circ \psi_*$ is the identity on ${\mathcal O}_F$,
from which it follows that $\psi_*$ is injective, and hence an isomorphism:
if $0 \le i \le m-1$, then
\[
\phi_*\circ\psi_*(V_{1,i})
= \phi_*(\psi((1,0),\delta_i))
= \phi_*(i_E((1,0),\delta_i))
= U_{1,i} = V_{1,i};
\]
if also $0 \le j \le n-1$, then
\begin{align*}
\phi_*\circ\psi_*(V_{2,in + j})
& = \phi_*(\psi((0,1),\delta_{in + j})) \\
& = \phi_*(i_E((1,1),\delta_{in + j}))
= U_{1,i}U_{2,j} = V_{2,in + j}.
\end{align*}
\end{proof}
\begin{proof}[Proof of Theorem~\ref{theorem:N2}]
By Proposition~\ref{prop:construct reps},
${\mathcal O}_{E(1,n)}$ is the universal $C^*$\nobreakdash- al\-ge\-bra for
collections $\{U_{1,0}\} \cup \{U_{2,0}, \dots, U_{2,n-1}\}$
of isometries satisfying
\[
U_{1,0}U_{1,0}^* = 1,
\quad
\sum_{i=0}^{n-1} U_{2,i}U_{2,i}^* = 1,
\quad\text{and}\quad
U_{1,0}U_{2,i} = U_{2,i}U_{1,0};
\]
i.e., ${\mathcal O}_{E(1,n)} \cong {\mathcal O}_n\otimes C(\field{T})$.
Suppose $m,n\ge 2$. If $\log_m n$ is irrational,
then $d: (a,b)\in\field{N}^2 \mapsto m^an^b$ is injective,
and ${\mathcal O}_{E(m,n)}$ is simple by Theorem~\ref{theorem:lexicographic}.
The existence and uniqueness of $l$ is elementary when $\log_m n$ is rational,
and repeated applications of Proposition~\ref{prop:factor}
give ${\mathcal O}_{E(m,n)} \cong {\mathcal O}_{E(l,1)}$.
\end{proof}
\section{Nuclearity}
Following \cite{dinhjfa,laca}, we call a cross section
$u:s\in P\mapsto u_s\in E_s\setminus\{0\}$
a {\em twisted unit\/} of $E$ if $u_su_t \in \field{C} u_{st}$ for every $s,t\in P$.
\begin{theorem}\label{theorem:nuclearity}
Let $G$ be a countable abelian group,
let $P$ be a subsemigroup of $G$ which contains the identity,
and let $E$ be a product system over $P$ of finite-dimensional Hilbert spaces.
If $E$ admits a twisted unit, then ${\mathcal O}_E$ is nuclear.
\end{theorem}
Our proof is modelled on the one given by Laca in \cite[Section~3]{laca}:
we realize ${\mathcal O}_E$ as a twisted semigroup crossed product of an AF algebra
by $P$, and deduce nuclearity from a theorem of Murphy
\cite[Theorem~3.1]{murphy}.
We begin by recalling the definition of a twisted semigroup crossed product.
Let $P$ be as in the theorem,
let $\beta$ be an action of $P$ as endomorphisms
of a unital $C^*$\nobreakdash- algebra $A$,
and let $\omega$ be a multiplier of $P$.
We call $(A,P,\beta,\omega)$ a {\em twisted semigroup dynamical system\/}.
Suppose $B$ is a unital $C^*$\nobreakdash- algebra.
An {\em isometric $\omega$\nobreakdash- representation\/}
of $P$ in $B$ is a map $V:P\to B$
such that each $V_s$ is an isometry and $V_sV_t = \omega(s,t)V_{st}$
for all $s,t\in P$.
A {\em covariant representation\/} of $(A,P,\beta,\omega)$ in $B$
is a pair $(\pi,V)$ consisting of a unital homomorphism $\pi:A\to B$
and an isometric $\omega$\nobreakdash- representation $V:P\to B$ such that
\[
\pi(\beta_s(a)) = V_s\pi(a)V_s^*
\qquad\text{for all $s\in P$ and $a\in A$.}
\]
A {\em crossed product\/} for $(A,P,\beta,\omega)$ is a triple
$(C,i_A,i_P)$ consisting of a unital $C^*$\nobreakdash- algebra $C$
and a covariant representation $(i_A,i_P)$ of $(A,P,\beta,\omega)$ in $C$
such that
(a) for every covariant representation $(\pi,V)$ of $(A,P,\beta,\omega)$
in a unital $C^*$\nobreakdash- algebra $B$,
there is a homomorphism $\pi\times V:C\to B$ such that
$(\pi\times V)\circ i_A = \pi$ and $(\pi\times V)\circ i_P = V$; and
(b) $C$ is generated as a $C^*$\nobreakdash- algebra by $i_A(A)\cup i_P(P)$.
\noindent The triple $(C,i_A,i_P)$ is unique up to canonical isomorphism.
\begin{proof}[Proof of Theorem~\ref{theorem:nuclearity}]
Let $u$ be a twisted unit for $E$.
By replacing $u_s$ with $\lVert u_s \rVert^{-1}u_s$,
we can assume that each $u_s$ is a unit vector.
Then $u_su_t = \omega(s,t)u_{st}$ determines a multiplier $\omega$ of $P$.
As in Section~\ref{section:abelian},
let ${\mathcal F}_E$ be the inductive limit $\varinjlim{\mathcal K}(E_s)$
under the embeddings $S\in{\mathcal K}(E_s) \mapsto S\otimes 1^t\in{\mathcal K}(E_{st})$,
and let $\iota_s$ be the canonical embedding of ${\mathcal K}(E_s)$ in ${\mathcal F}_E$.
Since tensoring on the left
by the rank-one projection $\rankone{u_r}{u_r}$
commutes with tensoring on the right by the identity,
for each $r\in P$
there is an endomorphism $\beta_r$ of ${\mathcal F}_E$ which satisfies
\[
\beta_r(\iota_s(S)) = \iota_{rs}((\rankone{u_r}{u_r})\otimes S)
\qquad\text{for all $s\in P$ and $S\in{\mathcal K}(E_s)$.}
\]
Note that $\beta_r$ is injective.
Moreover, for any $a\in {\mathcal F}_E$ we have
\begin{align*}
\beta_r\circ\beta_s(a)
& = (\rankone{u_r}{u_r}) \otimes (\rankone{u_s}{u_s}) \otimes a
= (\rankone{u_ru_s}{u_ru_s}) \otimes a \\
& = (\rankone{\omega(r,s)u_{rs}}{\omega(r,s)u_{rs}}) \otimes a
= (\rankone{u_{rs}}{u_{rs}}) \otimes a
= \beta_{rs}(a),
\end{align*}
and hence $\beta:P\to\End {\mathcal F}_E$ is a semigroup homomorphism.
Let $i_{\mathcal F}:{\mathcal F}_E\to{\mathcal O}_E$ be the embedding of
Proposition~\ref{prop:FE},
and define $i_P:P\to{\mathcal O}_E$ by $i_P(s) := i_E(u_s)$.
We claim that $({\mathcal O}_E,i_{\mathcal F},i_P)$ is a crossed product for the twisted
semigroup dynamical system $({\mathcal F}_E,P,\beta,\omega)$.
Each $i_P(s)$ is an isometry,
and
$i_P(s)i_P(t) = i_E(u_su_t) = \omega(s,t)i_E(u_{st}) = \omega(s,t)i_P(st)$,
so $i_P$ is an isometric $\omega$-representation of $P$ in ${\mathcal O}_E$.
If $x,y\in E_s$, then
\begin{align*}
i_{\mathcal F}(\beta_r(\iota_s(\rankone xy)))
& = i_{\mathcal F}(\iota_{rs}((\rankone{u_r}{u_r})\otimes(\rankone xy)))
= i_{\mathcal F}(\iota_{rs}(\rankone{u_rx}{u_ry})) \\
& = i_E(u_rx)i_E(u_ry)^*
= i_P(r)i_{\mathcal F}(\iota_s(\rankone xy))i_P(r)^*,
\end{align*}
and since elements of the form $\iota_s(\rankone xy)$ have dense
linear span in ${\mathcal F}_E$, we deduce that
$i_{\mathcal F}(\beta_r(a)) = i_P(r)i_{\mathcal F}(a)i_P(r)^*$
for all $r\in P$ and $a\in {\mathcal F}_E$.
Thus $(i_{\mathcal F},i_P)$ is a covariant representation of $({\mathcal F}_E,P,\beta,\omega)$
in ${\mathcal O}_E$.
We now verify condition (a) of a crossed product.
Suppose $(\pi,V)$ is a covariant representation of
$({\mathcal F}_E,P,\beta,\omega)$ in a unital $C^*$\nobreakdash- algebra $B$.
Define $\phi:E\to B$ by
\[
\phi(x) := \pi(\iota_{p(x)}(\rankone x{u_{p(x)}}))V_{p(x)}
\qquad\text{for $x\in E$.}
\]
We claim that $\phi$ is a Cuntz representation of $E$ in $B$.
If $x,y\in E_s$, then
\begin{align*}
\phi(y)^*\phi(x)
& = V_s^*\pi(\iota_s(\rankone y{u_s}))^*\pi(\iota_s(\rankone x{u_s}))V_s \\
& = V_s^*\pi(\iota_s((\rankone {u_s}y)(\rankone x{u_s})))V_s \\
& = (x\mid y)V_s^*\pi(\iota_s(\rankone{u_s}{u_s}))V_s \\
& = (x\mid y)V_s^*\pi(\beta_s(1))V_s = (x\mid y)1,
\end{align*}
and
\begin{equation}\label{eq:xystar}
\begin{split}
\phi(x)\phi(y)^*
& = \pi(\iota_s(\rankone x{u_s}))V_sV_s^*\pi(\iota_s(\rankone y{u_s}))^* \\
& = \pi(\iota_s((\rankone x{u_s})(\rankone{u_s}{u_s})(\rankone {u_s}y)))
= \pi(\iota_s(\rankone xy)).
\end{split}
\end{equation}
With ${\mathcal B}_s$ an orthonormal basis for $E_s$, \eqref{eq:xystar} gives
\[
\sum_{f\in {\mathcal B}_s} \phi(f)\phi(f)^*
= \sum_{f\in {\mathcal B}_s} \pi(\iota_s(\rankone ff))
= \pi(\iota_s(1)) = 1 = \phi(1).
\]
To see that $\phi$ is multiplicative,
suppose $x\in E_s$ and $y\in E_t$.
Then
\begin{align*}
\phi(x)\phi(y)
& = \pi(\iota_s(\rankone x{u_s}))V_s\pi(\iota_t(\rankone y{u_t}))V_t \\
& = \pi(\iota_s(\rankone x{u_s}))\pi(\beta_s(\iota_t(\rankone y{u_t})))V_sV_t \\
& = \pi(\iota_{st}((\rankone x{u_s})\otimes 1^t)
((\rankone{u_s}{u_s})\otimes(\rankone y{u_t})))V_sV_t \\
& = \pi(\iota_{st}((\rankone x{u_s})\otimes(\rankone y{u_t})))V_sV_t \\
& = \pi(\iota_{st}(\rankone{xy}{u_su_t}))V_sV_t \\
& = \pi(\iota_{st}(\rankone{xy}{u_{st}}))\overline{\omega(s,t)}V_sV_t \\
& = \pi(\iota_{st}(\rankone{xy}{u_{st}}))V_{st}
= \phi(xy).
\end{align*}
Thus $\phi$ is a Cuntz representation of $E$.
Let $\pi\times V:{\mathcal O}_E\to B$ be the integrated form of $\phi$;
that is, $\pi\times V$ satisfies
$(\pi\times V)\circ i_E = \phi$.
We claim that $(\pi\times V)\circ i_{\mathcal F} = \pi$
and $(\pi\times V)\circ i_P = V$, giving condition (a) of a crossed product.
If $x,y\in E_s$, then by \eqref{eq:xystar} we have
\[
(\pi\times V)\circ i_{\mathcal F}(\iota_s(\rankone xy))
= \pi\times V(i_E(x)i_E(y)^*)
= \phi(x)\phi(y)^*
= \pi(\iota_s(\rankone xy)),
\]
and since elements of the form $\iota_s(\rankone xy)$ have dense linear span
in ${\mathcal F}_E$, we deduce that $(\pi\times V)\circ i_{\mathcal F} = \pi$.
For the second part of the claim, we calculate
\begin{align*}
(\pi\times V)\circ i_P(s)
& = \pi\times V(i_E(u_s))
= \phi(u_s) \\
& = \pi(\iota_s(\rankone{u_s}{u_s}))V_s
= \pi(\beta_s(1))V_s = V_sV_s^*V_s = V_s.
\end{align*}
To verify condition (b) of a crossed product, suppose $x\in E_s$.
Then
\[
i_E(x) = i_E(x)i_E(u_s)^*i_E(u_s)
= i_{\mathcal F}(\iota_s(\rankone x{u_s}))i_P(s) \in C^*(i_{\mathcal F}({\mathcal F}_E)\cup i_P(P)),
\]
and since ${\mathcal O}_E$ is generated as a $C^*$\nobreakdash- algebra by $i_E(E)$,
this shows that it is also generated by $i_{\mathcal F}({\mathcal F}_E)\cup i_P(P)$, as required.
We have shown that $({\mathcal O}_E,i_{\mathcal F},i_P)$ is a crossed product for $({\mathcal F}_E,P,\beta,\omega)$.
Since ${\mathcal O}_E$ is the twisted semigroup crossed product
of a nuclear $C^*$\nobreakdash- algebra by an abelian semigroup of injective endomorphisms,
it is nuclear by \cite[Theorem~3.1]{murphy}.
\end{proof}
\begin{cor}\label{cor:nuclearity}
If $E$ is a twisted lexicographic product system over a subsemigroup
of a countable abelian group, then ${\mathcal O}_E$ is nuclear.
\end{cor}
\begin{proof} Since $u_s := (s,\delta_0)$ is a twisted unit for $E$,
Theorem~\ref{theorem:nuclearity} applies.
\end{proof}
\begin{remark} Suppose $P$ is a subsemigroup of a countable abelian group,
$d:P\to\field{N}^*$ is an injective homomorphism, and $\omega$ is a multiplier
of $P$. Let $E = E(d)^\omega$ be the lexicographic product system $E(d)$
twisted by $\omega$. Then ${\mathcal O}_E$ is a separable, unital
$C^*$\nobreakdash- algebra which is simple and purely infinite
(Theorem~\ref{theorem:main}) and nuclear (Corollary~\ref{cor:nuclearity}).
If $P = \field{N}^k$ and $\omega$ is trivial (i.e., $E = E(d)$),
then by the proof of Theorem~\ref{theorem:nuclearity}
${\mathcal O}_E$ is the crossed product of an AF algebra by $\field{N}^k$,
hence a full corner of the crossed product of an AF algebra by $\field{Z}^k$,
and hence belongs to the bootstrap class of algebras which satisfy
the Universal Coefficient Theorem \cite[Theorem~23.1.1]{blackadar}.
\end{remark}
|
\section{Introduction}
The bra-ket calculus invented by Dirac \cite{D} for unitary or Hilbert
spaces is one of the most coherent, efficient and elegant formalisms. This
is based on the natural relation among vectors and functionals represented
by ket and respectively bra vectors. In this way, one obtains simple
calculation rules which include the axioms of the scalar product as well as
the consequences of the Frechet-Riesz theorem. Grace of these qualities, the
Dirac formalism offers us the opportunity of working with general
operator relations, independent on the concrete representations that can
have very different features in the general case of the rigged Hilbert
spaces used in quantum theories.
However, in many problems the vector spaces can not be organized as unitary,
Euclidian or Hilbert spaces. We refer especially to the spaces of
finite-dimensional linear representations of the non-compact Lie groups
\cite{G,BR} where the bilinear forms we need to construct invariants are not
positive definite \cite{W}. In this case one uses the tensor calculus with
indices in upper and lower positions and bilinear (or inner) forms defined
with the help of a metric tensor. For unitary, Euclidian or
Hilbert spaces the tensor calculus is equivalent with the Dirac formalism in
a representation given by an orthonormal basis but for the spaces with
indefinite metric \cite{B} we have not yet a suitable Dirac formalism which
should reproduce all the mechanisms of calculus with covariant and
contravariant indices. For this reason we would like to propose here
a generalization of the Dirac formalism to vector spaces with indefinite
metric. We restrict ourselves only to the finite-dimensional case when these
spaces are called semi-unitary or semi-Euclidian \cite{ON}.
The main problem here is the generalization of the mutual bra-ket relation to
spaces with indefinite metric. In our opinion, this problem can not be
solved using only one vector space and its dual space. We mean that
starting with a space considered as being of covariant vectors (with
contravariant components) and with its dual space of contravariant
functionals,
we can not relate covariant vectors with contravariant functionals in a
satisfactory bra-ket formalism. This could be achieved only by introducing
new
ingredients, namely contravariant vectors and covariant functionals, which
should allow one to relate vectors and functionals of same kind (covariant or
contravariant). Therefore, the generalization of the Dirac formalism requires
to doubly the number of vector spaces like in the theory of general tensors
(with simple and dotted indices) where one uses four vector spaces associated
with the four unequivalent fundamental representations of the group of general
linear transformations \cite{W}. This means that we have already the framework
we need. It remains only to correctly define mutual bra-ket relations in
accordance with the mentioned exigency of a coherent bra-ket mechanism.
We show that this can be done with the help of two anti-linear
mappings that conserve the covariance. These will relate the spaces of
covariant and contravariant ket vectors with their corresponding spaces of bra
vectors. Moreover, we {\em couple} between themselves the spaces of ket vectors
as well as those of bra vectors through an isometry compatible with
the considered anti-linear mappings. In these conditions we can use the operator
of this isometry as {\em metric operator}. This will replace the metric tensor
taking over its role in defining scalar products.
Our objective is to generalize the Dirac bra-ket formalism in this mathematical
context which combines the framework of the general tensor calculus with the
theory of bilinear forms given by metric operators \cite{N}. We establish the
basic calculation rules in suitable notations and we verify that these lead to
the common tensor calculus in any representation given by a system of dual
bases. Actually, we shall see that our formalism recovers all the main results
of the theory of the general linear transformations but expressed only in terms
of components with usual covariant or contravariant indices. This is because
in the bra-ket formalism, where the complex conjugated components appear
naturally as the components of bra vectors, the artifice of dotted
indices is no more needed. On the other hand, we show that in
our approach we have all the technical advantages of the standard bra-ket
formalism. One of them is that we can work directly with general relations
involving operators instead of their matrix elements in particular
representations. Moreover, we can manipulate simultaneously different
spectral representations of these operators allowing
us to easily study the theory of finite-dimensional representation of the
non-compact Lie groups, including those of the symmetry group of the metric
operator.
The first step is to precise the mathematical framework of our attempt. This
is presented in the second section where we define the mutually related
pairs of coupled vector spaces which allows us to introduce the metric
operator and correct relations among vectors and functionals. The next
section is devoted to the specific symbols and notations we propose for
vectors and operators. In section 4 we construct two kind of compatible
hermitian forms, called dual forms and scalar products respectively, while
in section 5 we define the hermitian and Dirac conjugations for all the
linear operators we work. Section 6 is devoted to the theory of orthogonal
projection operators which help us to define pairs of coupled subspaces with
hermitian metric operators. The matrix representations in systems of dual
bases are studied in section 7 pointing out the advantages of the orthonormal
ones. The theory of basis transformations is briefly treated in section 8
where we give the form of the general linear transformations and we study
the symmetry transformations that leave invariant the form of the scalar
product. The example we propose in section 9 is the theory of
finite-dimensional representations of the $SL(2, \Comp)$ group with invariant
scalar products.
\section{Coupled vector spaces}
Our construction is based on a pair of complex vector spaces,
${\cal V}$ and $\hat{\cal V}$, of {\em covariant} vectors, $x,y,... \in
{\cal V}$, and
{\em contravariant} vectors, $\hat x, \hat y,...\in \hat{\cal V}$, respectively.
The dual of ${\cal V}$, denoted by $\overline{\cal V}$, is the space
of contravariant functionals, $\bar x,...$, while the dual of $\hat{\cal V}$,
denoted by $\overline{\hat{\cal V}}$, is the space of covariant functionals,
$\bar{\hat x}...$. The values of these functionals are
$\bar x(y)$ or $\bar{\hat x}(\hat y)$.
The basis vectors and the vector components are labeled by Latin indices
$i,j,k,...$ while the first Latin ones, $a,b,...$ are held for current needs.
We suppose that all the spaces we work are finite-dimensional of
dimension $N$ such that $i,j,...=1,2,...,N$. Furthermore, we consider
the covariant basis $\{e_{i}\}\subset {\cal V}$, the contravariant
basis $\{\hat e^{i}\}\subset \hat{\cal V}$, and the {\em canonical dual} bases,
from which
$\{\bar{e}^{i}\}\subset \overline{\cal V}$ is the contravariant one while
$\{\bar{\hat e_{i}}\}\subset \overline{\hat{\cal V}}$ is the covariant one.
Since we assume that these bases satisfy the usual duality conditions,
\begin{equation}\label{(dual)}
\bar{e}^{i}(e_{j})=\delta^{i}_{j}\,,\qquad
\bar{\hat e}_{i}(\hat e^{j})=\delta^{j}_{i}\,,
\end{equation}
we say that these form a {\em system of dual bases}.
A coherent bra-ket mechanism requires to {\em mutually relate} among
themselves the vectors and functionals. To this end, we introduce the
anti-linear mappings which conserve the covariance,
$\phi : {\cal V}\to \overline{\hat{\cal V}}$ and $\hat\phi
: \hat{\cal V}\to \overline{\cal V}$, defined by
\begin{equation}\label{(phi)}
\phi[e_{i}]=\bar{\hat e}_{i}\,,\quad
\hat \phi[\hat e^{i}]= \bar{e}^{i}\,,
\end{equation}
such that
\begin{equation}\label{(phixy1)}
\phi[x]=(x^{i})^{*}\bar{\hat e}_{i}\,,\quad
\hat \phi[\hat y]= (\hat y_{i})^{*}\bar{e}^{i}\,,
\end{equation}
for any $x=e_{i}x^{i}\in {\cal V}$ and $\hat y=\hat e^{i}\hat y_{i} \in
\hat{\cal V}$. Then from (\ref{(dual)}) we find
\begin{equation}\label{(phixy)}
\phi[x](\hat y)=(x^{i})^{*}\hat y_{i}=\left(\hat\phi[\hat y](x)\right)^{*}\,.
\end{equation}
In other respects, equations (\ref{(phixy1)}) show that
$\overline{\hat{\cal V}}\sim {\cal V}^{*}$ and
$\hat{\cal V}\sim \overline{{\cal V}^{*}}$ where ${\cal V}^{*}$
is the complex conjugate vector space of ${\cal V}$. Therefore, our
system of mutually related vector spaces is similar to that of the theory of
general tensors,
$({\cal V}, {\cal V}^{*}, \overline{{\cal V}}, \overline{{\cal V}^{*}})$.
Let us consider now the isomorphism $\eta : {\cal V}\to \hat{\cal V}$,
defined by the {\em invertible} matrix $|\eta|$ as
\begin{equation}\label{(etae)}
\eta\,e_{i}=\hat e^{j}\eta_{ji}.
\end{equation}
This {\em couples} each covariant vector $x=e_{i}x^{i}$ with the contravariant
vector $\hat x=\eta\, x =\hat e^{i} x_{i}$ of components $x_{i}= \eta_{ij}
x^{j}$. The corresponding isomorphism of the dual spaces, $\bar\eta:
\overline{\cal V}\to \overline{\hat{\cal V}}$, is given by
\begin{equation}
\bar{\eta}\,\bar{e}^{k}=(\eta^{-1})^{kj}\bar{\hat e}_{j}
\end{equation}
so that $\bar\eta\,\bar y(\eta\,x)= y(x)$.
Moreover, it is not difficult to verify that the isomorphism $\eta$
is compatible with the mappings $\phi$ and $\hat\phi$ (closing the diagram)
only if
\begin{equation}\label{(comp)}
\phi[x]=\bar\eta\,\hat\phi[\eta\,x], \quad \forall x\in {\cal V}.
\end{equation}
\begin{theor}
The condition {\rm (\ref{(comp)})} is accomplished if and only if the
matrix $|\eta|$ is hermitian, i.e. $\eta_{ij}=(\eta_{ji})^{*}$.
\end{theor}
\begin{demo}
Let us take $x=e_{i}$ and calculate $\bar{\hat e}_{i}=
\bar\eta\,\hat\phi[\eta\,e_{i}]$.
According to (\ref{(etae)}) and (\ref{(phi)}), this gives
$\bar{\hat e}_{i}=(\eta_{ji})^{*}\bar\eta\,\hat\phi[\hat e^{j}]
=(\eta^{-1})^{jk}(\eta_{ji})^{*}\bar{\hat e}_{k}$ from which we obtain the
desired result.
\end{demo}
In what follows we consider only invertible operators $\eta$ with hermitian
matrices in a given system of dual bases. They will be used as the {\em metric
operators} that define the {\em bilinear forms} of the spaces
${\cal V}$ and $\hat{\cal V}$,
\begin{eqnarray}\label{(bilf)}
h(x,y)&=&\phi[x](\eta\,y),\quad ~~~ x,y \in {\cal V},\\
\hat h(\hat x,\hat y)&=&\hat\phi[\hat x](\eta^{-1}\,\hat y),\quad
\hat x,\, \hat y \in \hat{\cal V}.
\end{eqnarray}
A little calculation giving us the bilinear forms in terms of vector
components points out that these are {\em hermitian} since
\begin{equation}\label{(herm)}
h(x,y)=h(y,x)^{*}\,,\quad
\hat h(\hat x,\hat y)=\hat h(\hat y,\hat x)^{*}.
\end{equation}
In addition, we can verify that
\begin{equation}\label{(izom)}
\hat h(\eta\,x,\, \eta\,y)=h(x,y)
\end{equation}
which means that the isomorphism $\eta$ is in fact an {\em isometry}
when $|\eta|=|\eta|^{+}$.
\begin{defin}
The spaces ${\cal V}$ and $\hat{\cal V}$ isometric through the metric operator
$\eta$ represent a pair of coupled vector spaces {\rm (cvs)} denoted by
$({\cal V},\hat{\cal V},\eta)$. The corresponding dual cvs are
$(\overline{\cal V},\overline{\hat{\cal V}},\bar\eta)$. We say that the
mappings $\phi$ and $\hat\phi$ mutually relate these pairs of cvs.
\end{defin}
\section{Notations}
The above defined related pairs of cvs represent the appropriate framework
of our generalized Dirac formalism. Let us start with the notations of the
basic elements.
\begin{defin}
The spaces ${\cal K}\equiv {\cal V}$ and $\hat{\cal K}\equiv \hat{\cal V}$
are the spaces of {\em ket-down} vectors, $\kd{~} \in {\cal K}$, or
of {\em ket-up} vectors, $\ku{~}\in \hat{\cal K}$.
\end{defin}
Thus in our formalism the covariant vectors appear as ket-down vectors while
the contravariant ones as ket-up vectors. The bra vectors mutually related with
these ket vectors can be defined with the help of the mappings $\phi$ and
$\hat \phi$.
\begin{defin}
The bra vector related to the ket-down vector $\kd{x}$ is the {\em bra-down}
vector $\bd{x}= \phi[\kd{x}] \in {\cal B}\equiv \overline{\hat{\cal V}}$
while the bra vector related to the ket-up vector $\ku{\hat x}$ is the
{\em bra-up} vector $\bu{\hat x} = \hat\phi[\ku{\hat x}] \in \hat{\cal B}
\equiv \overline{\cal V}$.
\end{defin}
In this manner we have related the covariant ket vectors of ${\cal K}$
with the covariant bra vectors of ${\cal B}$, which are just the
covariant functionals defined on the other ket space, $\hat{\cal K}$.
Similarly, the spaces of contravariant vectors, $\hat{\cal K}$ and
$\hat{\cal B}$, are also related between themselves even though the
contravariant functionals of $\hat{\cal B}$ are defined on ${\cal K}$.
Of course, the mutually related ket and bra vectors will be denoted
systematically with the same symbol as in the usual bra-ket formalism.
Apparently these crossed bra-ket relations seem to be forced but this is the
unique way to obtain well-defined hermitian bilinear forms compatible with
the natural duality. In other respects, our bra-ket relations are correct in
the sense that the bra vector related with a linear combination of ket
vectors is the corresponding anti-linear combination of bra vectors.
Obviously, this is because the mappings $\phi$ and $\hat\phi$ are anti-linear
(e.g. $\phi[\alpha\kd{x}+\beta\kd{y}]=\alpha^{*}\bd{x}+\beta^{*}\bd{y}$).
Let us consider now the linear operators defined on our cvs. In general, we
denote by $L({\cal V},{\cal V}')$ the set of the linear operators which map
${\cal V}$ onto ${\cal V}'$. In our case, the main pieces are the algebras
$L({\cal K},{\cal K}))$ and $L(\hat{\cal K},\hat{\cal K})$, but we are also interested by the operators which map
the spaces ${\cal K}$ and $\hat{\cal K}$ to each other. We start with the
observation that it is natural to denote by $/\!\!/\in L({\cal K},{\cal K})$ and
$\backslash \!\!\backslash\in L(\hat{\cal K},\hat{\cal K})$ the {\em identity} operators of these algebras since they
act upon the
ket-down and respectively ket-up vectors. Moreover, this notation helps us
to indicate the action of the operators of different kind by writing them
between identity operators. Thus, the operators of $L({\cal K},{\cal K})$ can be denoted
either simply by $A,B,...$ or by $\opd{A}, \opd{B},...$ in order to avoid
possible confusions with the operators of the other algebra,
$\opu{\hat A},\opu{\hat B},...\in L(\hat{\cal K},\hat{\cal K})$, or with those from
$L({\cal K},\hat{\cal K})$ or $L(\hat{\cal K},{\cal K})$ that have to be
delimited by both identity operators in a suitable order. On the other hand,
the notation we propose has the advantage of indicating the allowed algebraic
operations. For example, it is clear that the operators
$\backslash A/\in L({\cal K},\hat{\cal K})$ and
$/ B\backslash\in L(\hat{\cal K},{\cal K})$ can be multiplied to each other
while their sum does not make sense.
A special case is that of the metric operators, $\eta\in L({\cal K},
\hat{\cal K})$ and $\eta^{-1}\in L(\hat{\cal K},{\cal K})$, which play the
central role in our construction. They will be represented by
\begin{equation}
\eta\equiv \backslash\!/ \,,\quad \eta^{-1}\equiv /\!\backslash \,,
\end{equation}
in order to obtain the intuitive calculation rules
\begin{equation}
/\!\backslash \backslash\!/ = /\!\!/ \,,\quad \backslash\!//\!\backslash=\backslash \!\!\backslash\,.
\end{equation}
With these symbols, the pair of cvs of ket vectors can be denoted now by
$({\cal K}, \hat{\cal K}, \backslash\!/)$ while the related pair of bra cvs is
$(\hat{\cal B},{\cal B},/\!\backslash)$. The isometry $\backslash\!/ : {\cal K}\to \hat{\cal K}$
couples not only the vectors $\kd{x}$ and $\ku{\hat x}=\ku{\eta x}\equiv
\backslash\!/\!\!\kd{x}$, but also couples the
operators $\opd{A}\in L({\cal K},{\cal K})$ and $\opu{\hat A}\in L(\hat{\cal K},\hat{\cal K})$ through
\begin{equation}\label{(asop)}
\opu{\hat A}= \backslash\!/ A /\!\backslash \,.
\end{equation}
\section{Hermitian forms}
Here we can define two kind of brackets. The first one represents the values
of covariant or contravariant functionals.
We denote by $\bkud{\hat x}{y}\equiv \hat\phi[\hat x](y)$ the value of the
contravariant functional $\bu{\hat x}$ calculated for the vector $\kd{y}$,
and by $\bkdu{y}{\hat x}\equiv\phi[y](\hat x)$,
the value of the covariant functional $\bd{y}$
calculated for $\ku{\hat x}$.
We say that the mappings
$\bkud{~}{~}: \hat{\cal K}\times {\cal K}\to \Comp$ and
$\bkdu{~}{~}: {\cal K}\times \hat{\cal K}\to \Comp$ are {\em dual
forms}.
From (\ref{(phixy)}) it results that the dual forms are {\em hermitian}, i.e.
\begin{equation}\label{(bk1)}
\bkdu{y}{\hat x}=\bkud{\hat x}{y}^{*}\,, \quad \forall\,
\kd{y}\in {\cal K},\,\ku{\hat x}\in \hat{\cal K}.
\end{equation}
The second kind of brackets are just the hermitian bilinear forms defined by
(\ref{(bilf)}). Since the metric operator is invertible, these bilinear forms
are nondegenerate and, therefore, can be called {\em scalar products}
\cite{ON}. In our new notation the scalar product of ${\cal K}$,
$\bkdd{~}{~}: {\cal K}\times {\cal K}\to \Comp$, has the values
\begin{equation}
\bkdd{x}{y}=
\bd{x}\!\!\backslash\!/\!\!\kd{y}\equiv h(x,y)\,, \quad \kd{x},\kd{y}\in {\cal K},
\end{equation}
while that of $\hat{\cal K}$, $\bkuu{~}{~}: \hat{\cal K}\times\hat{\cal K}\to
\Comp$, gives
\begin{equation}
\bkuu{\hat x}{\hat y}= \bu{\hat x}\!\!/\!\backslash\!\!\ku{\hat y}
\equiv \hat h(\hat x,\hat y)\,, \quad \ku{\hat x},\ku{\hat y}\in \hat{\cal K}\,.
\end{equation}
The equations (\ref{(herm)}) which show that these scalar products are hermitian
take the form
\begin{equation}
\bkdd{x}{y}=\bkdd{y}{x}^{*}\,,\quad \bkuu{\hat x}{\hat y}=
\bkuu{\hat y}{\hat x}^{*}\,.
\end{equation}\label{(izom1)}
Other useful relations can be written starting with the coupled vectors
$\ku{\hat x}=\backslash\!/\!\!\kd{x}$ and $\ku{\hat y}=\backslash\!/\!\!\kd{y}$. Thus we can find
equivalences among the values of the dual forms and those of the scalar
products (e.g. $\bkdu{x}{\hat y}=\bkdd{x}{y}$, $\bkud{\hat x}{y}=\bkuu{\hat x}
{\hat y}$, etc.) or to recover equation (\ref{(izom)}) giving us the
isometry of cvs,
\begin{equation}\label{(xyxy)}
\bkdd{x}{y}=\bkuu{\hat x}{\hat y}\,.
\end{equation}
All these brackets can be imagined as resulting from the traditional
``juxtaposition" of the Dirac formalism \cite{D}. For example, we can write
$\bkdu{y}{\hat x}=(\bd{y})(\ku{\hat x})$, $\bkdd{x}{y}=\bd{x}
(\backslash\!/\!\!\kd{y})=(\bd{x}\!\!\backslash\!/)\kd{y}$, and so on. Hence the conclusion is that
we can combine ket and bra vectors of any kind in order to write brackets. If
the slash-lines are parallel we obtain a dual form but when these are not parallel,
leaving an empty angle, then we understand that therein is a metric operator
giving us a scalar product.
The orthogonality is defined by the scalar product which play the same
role as those of unitary or Hilbert spaces with the difference that
here the ``squared norm" $\bkdd{x}{x}$ can take any real
value, including $0$ even for vectors $\kd{x}\not=0$.
From (\ref{(xyxy)}) we see that if two ket-down vectors are orthogonal then
their coupled ket-up vectors are also orthogonal. As mentioned before, the cvs
are isometric in the sense that for two coupled ket vectors we have
$\bkdd{x}{x}=\bkuu{\hat x}{\hat x}$. Consequently, the separation of the orbits
for which this number is positive, negative or zero, can be done simultaneously
for both cvs \cite{B}.
\section{Hermitian and Dirac conjugations}
Now we have all the elements for defining the hermitian
conjugation that gives us the {\em hermitian adjoint} operators of our
linear operators. First we consider the operators from
$L({\cal K},\hat{\cal K})$ and $L(\hat{\cal K},{\cal K})$ and we define:
\begin{defin}
The hermitian adjoint operators of $\backslash A/ \in L({\cal K}, \hat{\cal K})$
and $/ B\backslash \in L(\hat{\cal K}, {\cal K})$ are
$\backslash A^{+}/ \in L({\cal K}, \hat{\cal K})$
and $/ B^{+}\backslash \in L(\hat{\cal K}, {\cal K})$
which satisfy
\begin{eqnarray}
\bd{x}\!A^{+}\!\kd{y}=\bd{y}\!A\!\kd{x}^{*}\,,\quad \forall\, \kd{x}\,,\kd{y}\in
{\cal K}\,,\\
\bu{\hat x}\!B^{+}\!\ku{\hat y}=\bu{\hat y}\!B\!\ku{\hat x}^{*}
\,,\quad \forall\, \ku{\hat x}\,,\ku{\hat y}\in \hat{\cal K}\,.
\end{eqnarray}
If $A=A^{+}$ or $B=B^{+}$ we say that these operators are {\em hermitian}.
\end{defin}
Obviously, from (\ref{(herm)}) we see that the metric operators are hermitian,
\begin{equation}
\backslash\!/^{+}=\backslash\!/\,,\quad /\!\backslash^{+}=/\!\backslash\,.
\end{equation}
For other operators the situation is more complicated as it results from
the following definitions.
\begin{defin}
The hermitian adjoint operator of $/A/\in L({\cal K},{\cal K})$ is the
operator $(\opd{A})^{+}=\opu{A^{+}} \in L(\hat{\cal K},\hat{\cal K})$ which
accomplishes
\begin{equation}\label{(herma)}
\bd{x}\!A^{+}\backslash\!/\!\! \kd{y}=\bd{y}\!\!\backslash\!/ A\!\kd{x}^{*}\,, \quad \forall
\kd{x},\kd{y}\in {\cal K}\,.
\end{equation}
For $\opu{B}\in L(\hat{\cal K},\hat{\cal K})$ the hermitian conjugation is defined by
\begin{equation}\label{(hermb)}
\bu{\hat x}\!B^{+}/\!\backslash\!\! \ku{\hat y}=
\bu{\hat y}\!\!/\!\backslash B\!\ku{\hat x}^{*}\,,
\quad \forall \ku{\hat x},\ku{\hat y}\in \hat{\cal K}\,,
\end{equation}
where $/ B^{+}/=(\backslash B\backslash)^{+}\in L({\cal K},{\cal K})$.
\end{defin}
One can convince ourselves that the mutually related bra vectors with
the ket vectors $/A\kd{x}$ and $\backslash B\ku{y}$ are $\phi[/A\kd{x}]=
\bd{x}A^{+}\backslash$ and $\hat \phi[\backslash B\ku{y}]=\bu{y}B^{+}/$ respectively.
\begin{defin}
We say that the operators which satisfy
\begin{equation}
\opu{A^{+}}=\backslash\!/ A/\!\backslash \,, \quad \opd{B^{+}}=/\!\backslash B\backslash\!/\,,
\end{equation}
are hermitian with respect to the metric $\backslash\!/$, or simply {\em semi-hermitian}.
\end{defin}
In other words an operator $/A/\in L({\cal K},{\cal K})$ is semi-hermitian if its adjoint
operator coincides with its coupled operator $\backslash \hat A\backslash$ defined by
(\ref{(asop)}).
We specify that these definitions can be formulated in terms of dual form.
For example, if we take $\ku{\hat y}=\backslash\!/\!\!\kd{y}$ then (\ref{(herma)}) and
(\ref{(hermb)}) can be rewritten as
\begin{eqnarray}
\bd{x}\!A^{+}\!\ku{\hat y}&=&\bu{\hat y}\!A\!\kd{x}^{*}\,,\quad
\forall\, \kd{x}\in {\cal K},\, \ku{\hat y}\in \hat{\cal K}\,,\label{(hermab)}\\
\bu{\hat x}\!B^{+}\!\kd{y}&=&\bd{y}\!B\!\ku{\hat x}^{*}\,,\quad
\forall\, \kd{y}\in {\cal K},\,\ku{\hat x}\in \hat{\cal K}.
\end{eqnarray}
Particularly, from (\ref{(bk1)}) we can draw the
conclusion that the identity operators are semi-hermitian, $(/\!\!/)^{+}=\backslash \!\!\backslash$.
In practice it is convenient to introduce another conjugation operation
which should unify the above definitions. This is just the generalization of
the familiar Dirac conjugation of the theory of four-component spinors.
\begin{defin}\label{(adjd)}
Given an operator $X$, the {\em Dirac adjoint} operator of $X$ is
\begin{equation}
\overline{X}=\left\{
\begin{array}{cll}
/\!\backslash X^{+}\backslash\!/& if &X\in L({\cal K},{\cal K})\,,\\
\backslash\!/ X^{+}/\!\backslash& if &X\in L(\hat{\cal K},\hat{\cal K})\,,\\
X^{+}& if &X\in L({\cal K},\hat{\cal K})\, or\, L(\hat{\cal K},{\cal K}) \,.
\end{array}\right.
\end{equation}
The operator $X$ is {\em self-adjoint} if $\,\overline{X}=X$.
\end{defin}
According to this definition all the hermitian and semi-hermitian operators
are self-adjoint, including the metric and the identity operators,
\begin{equation}
\overline{\backslash\!/}=\backslash\!/\,,\quad
\overline{/\!\backslash}=/\!\backslash\,,\quad
\overline{/\!\!/}=/\!\!/\,,\quad
\overline{\backslash \!\!\backslash}=\backslash \!\!\backslash\,.
\end{equation}
Furthermore, it is not difficult to demonstrate that
\begin{equation}
\overline{\left(\overline{X}\right)}=X\,,
\end{equation}
and that for two operators, $A$ and $B$, from the same linear space $L$ and
$\alpha, \beta \in \Comp$ we have
\begin{equation}
\overline{(\alpha A+\beta B)}=\alpha^{*}\,\overline{A}+\beta^{*}\,
\overline{B}\,.
\end{equation}
If the multiplication of two operators makes sense,
e.g. $A,\,B\in L({\cal K},{\cal K})$ or $A\in L({\cal K},\hat{\cal K})$
and $B\in L(\hat{\cal K},{\cal K})$, etc, then we can show that
\begin{equation}
\overline{AB}=\overline{B}\,\overline{A}\,.
\end{equation}
Thus we obtain a simple and homogeneous set of calculation rules for all the
linear operators we manipulate. However, despite of this advantage, we prefer
to use here the hermitian conjugation rather than the Dirac one since we are
interested to follow the coherence of the presented formalism in its all
details.
\section{Projection operators}
In the case of our cvs, where the orthogonality is defined by the metric
operator, the problem of the decomposition in orthogonal subspaces is more
complicated than that of unitary spaces \cite{ON} and, therefore, the theory of
projection operators needs some specifications.
Let us consider the cvs $({\cal K},\hat{\cal K},\backslash\!/)$ and an idempotent
operator $/P/$, satisfying $P^{2}=P$, coupled with $\backslash \hat P\backslash =
\backslash\!/ P/\!\backslash$ . Then $P$ is the projection operator on the subspace
$P{\cal K}\subset
{\cal K}$ while $\opu{\hat P}$, is the projection operator on the subspace
$\hat P\hat{\cal K}\subset \hat{\cal K}$. Two projection operators,
$/P_{1}/$ and
$/P_{2}/$, are {\em additive} if $P_{1}P_{2}= P_{2}P_{1}=0$ since
then $P_{1}+P_{2}$ is a projection operator too.
The operators $P_{a}, a=1,2,...,n$, which satisfy
\begin{equation}
P_{a}P_{b}=\delta_{a}^{b}P_{a}\,, \quad \not\!\!{\Sigma}\,,
\end{equation}
form a set of additive projection operators. This set is called
{\em complete} if
\begin{equation}
\sum_{a}P_{a}=/\!\!/\,.
\end{equation}
However, these projection operators do not have good orthogonality
properties since the additive ones generally are not orthogonal. For this
reason one prefers the term perp instead of orthogonal \cite{ON}.
\begin{defin}
Two projection operators, $P_{1}$ and $P_{2}$, are perp if they satisfy
\begin{equation}
P_{1}^{+}\backslash\!/ P_{2}=0
\end{equation}
Then the projection subspaces are called perp to each other.
\end{defin}
The projection subspaces of these projection operators can not be
coupled anytime because there is the risk to find that the restriction of
the metric operator to these subspaces is no more hermitian. Two
projection subspaces, $P{\cal K}$ and $\hat P\hat{\cal K}$, can be coupled
only if their metric operator $\hat P\backslash\!/ P$ is hermitian. This requires
$\hat P=P^{+}$ which means that $P$ must be semi-hermitian (self-adjoint).
Then this has the property
\begin{equation}\label{(vpr)}
P^{+}\backslash\!/ =\backslash\!/ P\,,
\end{equation}
which indicates that the subspaces $P{\cal K}$ and $P^{+}\hat{\cal K}$ are
{\em invariant} subspaces of the metric operators $\backslash\!/$ and $/\!\backslash$. In these
conditions the operators
\begin{equation}
\backslash\!/_{P}=P^{+}\backslash\!/ P= P^{+}\backslash\!/ = \backslash\!/ P\,,\quad
/\!\backslash_{P}=P/\!\backslash P^{+}= P/\!\backslash = /\!\backslash P^{+},
\end{equation}
can be considered as metric operators since they are hermitian and
invertible in the sense that
\begin{equation}
\backslash\!/_{P}/\!\backslash_{P}=P^{+}\,,\quad
/\!\backslash_{P}\backslash\!/_{P}=P\,.
\end{equation}
\begin{defin}
The invariant subspaces $P{\cal K}$ and $P^{+}\hat{\cal K}$ isometric
through $\backslash\!/_{P}$ represent a pair of coupled subspaces {\em (css)} denoted by
$(P{\cal K},P^{+}\hat{\cal K}, \backslash\!/_{P})
\subset ({\cal K}, \hat{\cal K}, \backslash\!/)$.
\end{defin}
The related css, $(\hat {\cal B}P, {\cal B}P^{+}, /\!\backslash_{P})$, can be
defined using the restrictions to $P{\cal K}$ and
$P^{+}\hat{\cal K}$ of the mappings $\phi$ and $\hat \phi$.
Now the theory of orthogonal decomposition can be done in terms of css
determined by semi-hermitian projection operators.
\begin{defin}
If two semi-hermitian projection operators are perp then they as well as their
projection css are called orthogonal.
\end{defin}
This definition is justified by the fact that the semi-hermitian projection
operators have similar properties as the usual hermitian ones.
\begin{theor}
Two semi-hermitian projection operators are orthogonal if and only if they are
additive.
\end{theor}
\begin{demo} Let us consider that $P_{1}$ and $P_{2}$ are semi-hermitian and
additive, satisfying $P_{1}P_{2}=0$. Then, according to (\ref{(vpr)}), we can
write $0=\backslash\!/ P_{1}P_{2}=P_{1}^{+}\backslash\!/ P_{2}$ which means that these projection
operators are orthogonal. From the same relation it results that the orthogonal
projection operators must be additive since, by hypothesis, $\backslash\!/$ is invertible.
\end{demo}\\
The consequence is that the css
$({\cal K}_{3},\hat {\cal K}_{3}, \backslash\!/_{3})$
defined by the projection operator $P_{3}=P_{1}+P_{2}$ is the {\em direct sum}
\begin{equation}
({\cal K}_{3},\hat {\cal K}_{3}, \backslash\!/_{3})=
({\cal K}_{1},\hat {\cal K}_{1}, \backslash\!/_{1})\oplus
({\cal K}_{2},\hat {\cal K}_{2}, \backslash\!/_{2})
\end{equation}
where
\begin{equation}
\backslash\!/_{3}=\backslash\!/ (P_{1}+P_{2})=\backslash\!/_{1}+\backslash\!/_{2}\,.
\end{equation}
If $P_{2}=/\!\!/-P_{1}$ then the css defined by $P_{1}$ and $P_{2}$ are {\em
orthogonal complements}. In general, a complete set of additive
projection operators is a complete set of orthogonal projection operators if
all of of these operators are semi-hermitian.
\section{Matrix representations}
\subsection{Dual bases}
According to the usual terminology, any system of dual bases defines
a {\em matrix representation} of the cvs of ket and bra vectors.
In our new notation the vectors of the dual bases we have introduced in
section 2 are
\begin{eqnarray}
\kd{(i)}\equiv e_{i}\in {\cal K} \,, &\quad&
\ku{(i)}\equiv \hat e^{i}\in \hat{\cal K}\,,\label{(b1)}\\
\bu{(i)}\equiv {\bar e}^{i} \in \hat{\cal B}\,, &\quad&
\bd{(i)}\equiv \bar{\hat e}_{i}\in {\cal B}\,.\label{(b2)}
\end{eqnarray}
We consider that the indices $\backslash(i)$ or $(i)/$ are in {\em upper} position
while $/(i)$ or $(i)\backslash$ are in {\em lower} position and we use the summation
convention over dummy indices in opposite positions (e.g. we sum over $i$ in
expressions where we find $... \backslash (i)...(i)\backslash...$ or $... / (i)...(i)
/...$).
In a given representation, the main tools of our formalism are the
duality conditions (\ref{(dual)}), written now as
\begin{equation}\label{(dc)}
\bkud{(i)}{(j)}=\delta^{i}_{j}\,,\quad
\bkdu{(i)}{(j)}=\delta_{i}^{j}\,,
\end{equation}
and the {\em completeness relations}
\begin{equation}\label{(cr)}
\kd{(i)}\bu{(i)}=/\!\!/\,, \quad \ku{(i)}\bd{(i)}=\backslash \!\!\backslash\,,
\end{equation}
which show that the sets of {\em elementary} projection operators
\begin{equation}\label{(epr)}
/P_{i}/=\kd{(i)}\bu{(i)} \quad \not\!\!{\Sigma}\,,\quad i=1,2,...,N,
\end{equation}
and respectively $P_{i}^{+}, i=1,2,...,N$, are complete sets of additive
projection operators.
These formulas contain all the information concerning the vector calculus
with upper and lower indices, allowing us to express the final results in
terms of vector components and matrix elements. A special role play the
matrix elements of the metric operator,
\begin{eqnarray}
\bkdd{(i)}{(j)}&=&
\bd{(i)}\!\!\backslash\!/\!\!\kd{(j)}\equiv\eta_{ij}\,,\\
\bkuu{(i)}{(j)}&=&
\bu{(i)}\!\!/\!\backslash\!\!\ku{(j)}\equiv
{(\eta^{-1})}^{ij}\,,
\end{eqnarray}
which change the positions of indices. The components of ket-down vectors are
$\bkud{(i)}{x}\equiv x^{i}$ or $\bkdd{(i)}{x}\equiv x_{i}=\eta_{ij}x^{j}$,
and similarly for the ket-up vectors. The components of the corresponding bra
vectors have to be obtained through complex conjugation, according to the
usual properties of the dual forms or scalar products.
However, the most important is that now we are able to
explicitly use spectral representations. For example, the operator
$\opd{A}$ can be written as
\begin{equation}
\opd{A}=\kd{(i)}\bu{(i)}\!A\!\kd{(j)}\bu{(j)}=\kd{(i)}A^{i\,\cdot}_{\cdot\,j}
\bu{(j)}
\end{equation}
where $A^{i\,\cdot}_{\cdot\,j}$ are the matrix elements in usual notation.
The adjoint operator of $A$ is
\begin{equation}
\opu{A^{+}}=\ku{(i)}\bd{(i)}\!A^{+}\!\ku{(j)}\bd{(j)}=\ku{(i)}
(A_{\cdot\,i}^{j\,\cdot})^{*}\bd{(j)}
\end{equation}\label{(hermaa)}
since from (\ref{(hermab)}) we have
\begin{equation}\label{(hermaa)}
(A^{+})_{i\,\cdot}^{\cdot\,j}\equiv \bd{(i)}\!A^{+}\!\ku{(j)}=
\bu{(j)}\!A\!\kd{(i)}^{*}\equiv(A_{\cdot\,i}^{j\,\cdot})^{*}\,.
\end{equation}
If, in addition, $A$ is semi-hermitian then
$\bd{(i)}\!A^{+}\!\ku{(j)}=\bd{(i)}\!\!\backslash\!/ A/\!\backslash\!\!\ku{(j)}$
and using (\ref{(hermaa)}) we recover the well-known property
$\eta_{ik}A^{k\,\cdot}_{\cdot\,l}(\eta^{-1})^{lj}=
(A_{\cdot\,i}^{j\,\cdot})^{*}$.
In any system of dual bases the traces of the operators $A\in L({\cal K},{\cal K})$ and
$B\in L(\hat{\cal K},\hat{\cal K})$ are defined by
\begin{equation}
{\rm Tr}(A)=\bu{(k)}\!A\!\kd{(k)}\equiv A^{k\,\cdot}_{\cdot \, k}\,,\quad
{\rm Tr}(B)=\bd{(k)}\!B\!\ku{(k)}\equiv B_{k\,\cdot}^{\cdot \, k}\,.
\end{equation}
It is easy to show that the coupled operators (\ref{(asop)}) have the same
trace.
\subsection{Orthonormal bases}
In applications one prefers the representations given by the {\em orthonormal}
dual bases where the matrix of the metric operator is diagonal,
\begin{equation}\label{(etad)}
\eta_{ij}\equiv\bkdd{(i)}{(j)}=\eta_{i}\delta_{ij} \quad \not\!\!{\Sigma}\,,\quad
\eta_{i}=\pm1\,,
\end{equation}
since there the ``squared norms" have the simplest expressions,
\begin{equation}
\bkdd{x}{x}=\sum_{i=1}^{N}\eta_{i}|\bkud{(i)}{x}|^{2}\,.
\end{equation}
The numbers $\eta_{i}$ may take $n_{+}$ times the value $1$ and $n_{-}=N-n_{+}$
times the value $-1$. Thus in orthonormal bases the metric operator is defined
by its {\em signature} that can be given either explicitly as a sequence of
signs or simply as $(n_{+},n_{-})$. We note that $n_{-}$ is called the
{\em index} of the metric operator \cite{ON}.
The main advantage of the orthonormal bases is that there the elementary
projection operators (\ref{(epr)}) are semi-hermitian and, consequently, they
are orthogonal to each other. Of a particular interest are the orthogonal
projection operators
\begin{equation}
P_{+}=\sum_{\eta_{i}=1}P_{i}\,, \qquad
P_{-}=\sum_{\eta_{i}=-1}P_{i}\,,
\end{equation}
which satisfy $/P_{+}/+/P_{-}/=/\!\!/$.
They split the space ${\cal K}$ into the pair of unitary spaces
${\cal K}_{+}=P_{+}{\cal K}$ and ${\cal K}_{-}=P_{-}{\cal K}$, of dimensions
${\rm dim}\,{\cal K}_{+}=n_{+}$ and ${\rm dim}\,{\cal K}_{-}=n_{-}$. Since
the coupled space can be split in the same manner we have
\begin{equation}\label{(deco)}
({\cal K},\hat{\cal K},\backslash\!/)=
({\cal K}_{+},\hat{\cal K}_{+},\backslash\!/_{+})\oplus
({\cal K}_{-},\hat{\cal K}_{-},\backslash\!/_{-})
\end{equation}
where
\begin{equation}
\backslash\!/=\backslash\!/_{+} +\backslash\!/_{-}\,;\quad \backslash\!/_{+}=\backslash\!/ P_{+}\,,\quad \backslash\!/_{-}=\backslash\!/ P_{-}\,.
\end{equation}
Obviously, the same decomposition can be done for the related cvs of bra
vectors.
Finally we specify that in the particular case of the metric operators with
signature $(N,0)$ or $(0,N)$,
the matrix $|\eta|$ in orthonormal bases coincides up to sign with the unit
matrix. Then our scalar products become usual inner forms (i.e. scalar
products in the sense of the theory of Hilbert spaces) and, therefore,
${\cal K}$ and $\hat{\cal K}$ will be {\em unitary} spaces. In this
situations we have two options. The first one is to keep the cvs structure if
this is appropriate for our problem. The second option is to consider
only one ket space ${\cal K}\equiv \hat{\cal K}$ in usual Dirac formalism,
with $\ku{~}=\kd{~}=\ket{~}$ and $\backslash\!/=/\!\backslash=/\!\!/=\backslash \!\!\backslash=I$, where $I$ is the
identity operator on ${\cal K}$.
\section{Basis transformations}
\subsection{General linear transformations}
The change of the system of dual bases of our related pairs of cvs can be done
using for each basis an arbitrary linear transformation, but then it is
possible to obtain new bases
which do not satisfy the canonical duality conditions or giving a non-hermitian
matrix for the metric operator. In order to avoid these unwanted
situations, we consider only the transformations which preserve
the cvs structure in the sense that (i) leave invariant the duality conditions
and (ii) transform the hermitian matrix $|\eta|$ into another hermitian matrix,
$|\eta'|$.
\begin{theor}
The general form of a transformation which satisfies the conditions (i) and
(ii) is
\begin{eqnarray}
&&\kd{(i)}\to\kd{(i)'}=/ T\kd{(i)},\nonumber\\
&&\ku{(i)}\to\ku{(i)'}=\backslash (T^{-1})^{+}\ku{(i)},\label{(cant)}\\
&&\bu{(i)}\to \bu{(i)'}=\bu{(i)}T^{-1}/,\nonumber\\
&&\bd{(i)}\to\bd{(i)'}=\bd{(i)}T^{+}\backslash.\nonumber
\end{eqnarray}
where $T\in {\rm Aut}\,({\cal K})\subset L({\cal K},{\cal K})$.
\end{theor}
\begin{demo}
Let us start with the following linear transformation of the ket
vectors
\begin{equation}\label{(trket)}
\kd{(i)'}=/ T\kd{(i)}\,,\quad \ku{(i)'}=\backslash\tilde T\ku{(i)}\,,
\end{equation}
given by two operators $T\in {\rm Aut}({\cal K})$ and $\tilde T\in {\rm Aut}
(\hat{\cal K})$ arbitrarily chosen. The transformations of the bra vectors
which conserve the duality conditions are
\begin{equation}\label{(trbra)}
\bu{(i)'}=\bu{(i)}T^{-1}/\,,\quad \bd{(i)'}=\bd{(i)}{\tilde T}^{-1}\backslash\,.
\end{equation}
In these new bases the matrix of the metric operator remains hermitian only if
$\tilde T=(T^{-1})^{+}$ since then we have
\begin{equation}\label{(etap)}
\eta_{ij}'\equiv\bkdd{(i)'}{(j)'}=\bd{(i)}\! T^{+}\backslash\!/ T\!\kd{(j)}=
(T^{k\,\cdot}_{\cdot\,i})^{*}T^{l\,\cdot}_{\cdot\,j}\,\eta_{kl}\,.
\end{equation}
Consequently, the transformations (\ref{(trket)}) and (\ref{(trbra)}) take
the form (\ref{(cant)}).
\end{demo}
The conclusion is that $T$ is the operator of a {\em general linear}
transformation of the group $GL(N,\Comp)\subset L({\cal K},{\cal K})$. This means that
our formalism is equivalent with the representation theory of
general linear transformations in the vector spaces of the theory of
general tensors. Indeed, in a given system of dual bases, the components of the
ket-down and respectively ket-up vectors transform according to a pair of
unequivalent fundamental representations of the $GL(N,\Comp)$ group while
the bra components transform according to the corresponding complex conjugate
representations. Moreover, we observe that
the bra components of our formalism, $\bkdu{x}{(k)}\equiv (x^{k})^{*}$ and
$\bkud{\hat y}{(k)}\equiv (\hat y_{k})^{*}$, are just those replaced in the
theory of general tensors
by components carrying dotted indices in opposite positions. With this
specification, it is clear that the four transformations laws (\ref{(cant)})
correspond to the four unequivalent fundamental representations of the
$GL(N,\Comp)$ group \cite{W}.
The advantage of our formalism is that we work directly with the operators
$T$ instead of the matrices of their four representations. In this way we can
easily find the transformation laws of the matrices of all the operators we
use. The matrices of the operators $/A/$ and $\backslash B\backslash$ transform as
\begin{eqnarray}
A^{i\,\cdot}_{\cdot\,j}\to
A^{'i\,\cdot}_{\cdot\,j}\equiv \bu{(i)'}\!A\!\kd{(j)'}&=&
\bu{(i)}\!T^{-1}AT\!\kd{(j)}\nonumber\\
&=&(T^{-1})^{i\,\cdot}_{\cdot\,k}
A_{\cdot\,l}^{k\,\cdot}T^{l\,\cdot}_{\cdot\,j} \\
B_{i\,\cdot}^{\cdot\,j}\to
B_{i\,\cdot}^{'\cdot\,j}\equiv \bd{(i)'}\!B\!\ku{(j)'}&=&
\bd{(i)}\!T^{+}B(T^{-1})^{+}\!\ku{(j)}\nonumber\\
&=&(T_{\cdot\,i}^{k\,\cdot})^{*}
B^{\cdot\,l}_{k\,\cdot}\left[(T^{-1})^{j\,\cdot}_{\cdot\,l}\right]^{*}
\end{eqnarray}
while those of the operators from $L({\cal K},\hat{\cal K})$ and
$L(\hat{\cal K}, {\cal K})$ transform like the matrices of the metric
operators, according to equation (\ref{(etap)}).
These transformations leave invariant the traces of the operators
from $L({\cal K},{\cal K})$ and $L(\hat{\cal K},\hat{\cal K})$. Moreover,
one can show that any system of dual bases can be transformed at any time in
a system of othonormal bases using a suitable general transformation.
\subsection{Symmetry transformations}
The general transformations change the form of all the operator
matrices including that of the metric operator. However, there is a special
case of some transformations which do not change the matrix of the
metric operator.
\begin{defin}
The general transformations that leave invariant the matrix $|\eta|$
are called symmetry transformations.
\end{defin}
These transformations are of the form (\ref{(cant)}) but their operators
have special properties.
\begin{theor}
The operators $U$ of the symmetry transformations must satisfy the
{\em semi-unitarity} condition,
\begin{equation}\label{(etaun)}
\backslash U^{+}\backslash\!/ U/=\backslash\!/\,,
\end{equation}
which can be written as $\overline{U}=U^{-1}$.
\end{theor}
\begin{demo}
A transformation (\ref{(cant)}) generally change the matrix $|\eta|$
according to (\ref{(etap)}). This is a symmetry transformation only
if $\bkdd{(i)'}{(j)'}=\bkdd{(i)}{(j)}$. Hereby it results (\ref{(etaun)}).
\end{demo}\\
Consequently, the symmetry transformations have the form
\begin{eqnarray}
&&\kd{(i)}\to\kd{(i)'}=/ U\kd{(i)},\nonumber\\
&&\ku{(i)}\to\ku{(i)'}=\backslash \hat U\ku{(i)},\label{(cantu)}\\
&&\bu{(i)}\to \bu{(i)'}=\bu{(i)} U^{-1}/,\nonumber\\
&&\bd{(i)}\to\bd{(i)'}=\bd{(i)}\hat U^{-1}\backslash.\nonumber
\end{eqnarray}
where we recall that $\hat U=\backslash\!/ U/\!\backslash$ is the operator coupled with $U$.
The main virtue of these transformations is that they do not change the
expressions of scalar product in terms of vector components. In other words
these leave invariant not only the expressions of the dual forms but also those of
the scalar products. Moreover, when we work with orthonormal bases the symmetry
transformations conserve the orthogonality.
The semi-unitary operators $U$ which accomplish the condition
(\ref{(etaun)}) form a subgroup of $GL(N,\Comp)$, namely the {\em maximal
symmetry group} or the {\em gauge group} of the
metric operator $\backslash\!/$. Any system of dual bases defines a pair of
{\em coupled} fundamental representations of this group and its algebra
in the carrier spaces ${\cal K}$ and $\hat{\cal K}$.
Obviously, these representations are {\em equivalent} through the metric
operator.
A given pair of ket cvs can be used as carrier spaces for the
coupled {\em semi-unitary} representations
of any subgroup of the gauge group. In general, these
representations are reducible in usual sense but it is not sure that their
subspaces can be correctly coupled. For this reason we consider here a modified
definition of reducibility.
\begin{defin}\label{(redu)}
The coupled representations are irreducible
if their generators and the metric operator have no common non-trivial
invariant subspaces. Otherwise the representations are reducible.
\end{defin}
Since these representations are semi-unitary one can show that,
like in the unitary case, the reducible representations are decomposable (i.e.
fully reducible). Consequently, the original cvs can be written as a direct sum
of css carrying irreducible representations. Thus, using this definition we
preserve the important advantage of the css structure that guarantees the
existence of invariant scalar products.
\subsection{Group generators}
The properties of the $GL(N,\Comp)$ group and its subgroups are
well-studied but it is interesting to review few among them in our
formalism where we can work directly with the spectral representations of the
group generators.
Let us consider the related pairs of cvs carrying the fundamental
representations of the $GL(N,\Comp)$ group (\ref{(cant)}). In any system of
dual bases (\ref{(b1)}) and (\ref{(b2)}) the usual parametrisation of the
operators $T\in GL(N,\Comp)\subset L({\cal K},{\cal K})$ reads
\begin{equation}\label{(param)}
T(\omega)=e^{\omega^{ij}X_{ij}}\,, \quad X_{ij}=\kd{(i)}\bd{(j)}\!\!\backslash\!/,
\end{equation}
where $\omega^{ij}$ are arbitrary c-numbers and $X_{ij}$ are the ``real
generators" (having real matrix elements in orthonormal bases) which satisfy
\begin{equation}\label{(hermx)}
(X_{ij})^{+}=\backslash\!/ X_{ji}/\!\backslash \qquad \left(\overline{X}_{ij}=X_{ji}\right)\,.
\end{equation}
Hereby we can separate the $SL(N,\Comp)$ generators
\begin{equation}\label{(hij)}
H_{ij}=X_{ij}-\frac{1}{N}\eta_{ji}\,/\!\!/\,,
\end{equation}
which have the properties
\begin{equation}
\bkuu{(i)}{(j)}H_{ij}=0\,,\quad {\rm Tr}(H_{ij})=0\,.
\end{equation}
If we consider only the gauge group of the metric operator $\backslash\!/$ then we
have to use the same generators but with restrictions imposed upon
the parameter values. From equations (\ref{(etaun)}) and (\ref{(hermx)})
we obtain that the parameters of the gauge group must satisfy
the condition $\omega^{ij}+(\omega^{ji})^{*}=0$, which means that
\begin{equation}\label{(roio)}
\Re\, \omega^{ij}=-\Re\, \omega^{ji}\,,\quad \Im\,\omega^{ij}=
\Im\,\omega^{ji}\,.
\end{equation}
There are $N^{2}$ real parameters as it was expected since the gauge group of
the metric operator $\backslash\!/$ of signature $(n_{+},n_{-})$ is the
{\em semi-unitary} group $U(n_{+},n_{-})=U(1)\otimes SU(n_{+},n_{-})$.
When one uses the parameters $\omega ^{ij}$ then the operators (\ref{(hij)})
are considered as ``real" $SU(n_{+},n_{-})$ generators. However, the
canonical parametrisation with the {\em real} parameters (\ref{(roio)})
reads $\omega^{ij}H_{ij}=-i(\Re\omega^{ij}A_{ij}+\Im\omega^{ij}S_{ij})$
involving the {\em semi-hermitian} (self-adjoint) $SU(n_{+},n_{-})$
generators defined as
\begin{eqnarray}
A_{ij}&=&\frac{i}{2}(X_{ij}-X_{ji})+\frac{1}{N}\Im\,\eta_{ji}\,/\!\!/\,,\nonumber \\
S_{ij}&=&-\frac{1}{2}(X_{ij}+X_{ji})+\frac{1}{N}\Re\,\eta_{ji}\,/\!\!/\,.\label{(AS)}
\end{eqnarray}
It is clear that the antisymmetric ones, $A_{ij}$, are the generators of
the subgroup $SO(n_{+},n_{-})\subset SU(n_{+},n_{-})$ \cite{BR}.
In the particular case of real cvs the gauge group reduces
to $O(n_{+},n_{-})$. On the other hand, with the help of the generators
(\ref{(AS)}) one can show that the group $SL(N,\Comp)$ is the complexification
of the $SU(n_{+},n_{-})$ group in the sense that in a parametrisation with
real numbers the $SL(N,\Comp)$ generators are $A_{ij}$, $S_{ij}$, $iA_{ij}$
and $iS_{ij}$.
An interesting problem is how transform the generators $X_{ij}$
when we change the bases through a general linear transformation (\ref{(cant)}).
In our formalism it is easy to show that, according to (\ref{(param)}), the
transformed generators are
\begin{equation}
X'_{ij}=\kd{(i)'}\bd{(j)'}\!\!\backslash\!/=T X_{ij}/\!\backslash T^{+}\backslash\!/=TX_{ij}\overline{T}.
\end{equation}
Particularly, if we consider only symmetry transformations, $T=U$,
then from (\ref{(etaun)}) we recover the usual transformation law
of the $SU(n_{+},n_{-})$ generators,
\begin{equation}
H'_{ij}=U H_{ij}U^{-1}=U^{k\,\cdot}_{\cdot\,i}
(U^{l\,\cdot}_{\cdot\,j})^{*}H_{kl}\,,
\end{equation}
which indicates that they transform according to the {\em adjoint}
representation of $SU(n_{+},n_{-})$.
\section{The semi-unitary representations of the $SL(2,\Comp)$ group}
The application presented in order to illustrate how works our formalism is
the problem of the finite-dimensional representations of the $SL(2,\Comp)$
group with invariant scalar products. It is well-known that the
finite-dimensional irreducible representations of the $sl(2,\Comp)$ algebra
can be constructed with the help of those of the $su(2)$
algebra \cite{W}. However, in general, these do not have invariant scalar
products under $SL(2,\Comp)$ transformations. In practice these scalar
products are defined in each
particular case of physical interest separately starting with a suitable
representation which is often reducible. In this section we would like to
present the general theory of the semi-unitary and irreducible
finite-dimensional representations of $sl(2,\Comp)$, in coupled carrier
spaces where the invariant scalar products are well-defined.
\subsection{The representations of $su(2)$ in cvs}
The problem of the irreducible representations of $su(2)$ in cvs reduces to
that of the canonical irreducible representations in unitary spaces
${\cal K}^{j}\sim \hat{\cal K}^{j}$ of weight $j$ \cite{AB}. Therefore the
non-trivial
cvs may be carrier spaces only for reducible coupled representations. These cvs
have the general structure
\begin{equation}\label{(cvsrr)}
({\cal K},\hat {\cal K}, \backslash\!/)=\sum_{j\in {\bf J}}\oplus
({\cal K}^{j},\hat {\cal K}^{j}, \backslash\!/^{j})
\end{equation}
where
\begin{equation}
\backslash\!/=\sum_{j\in {\bf J}} \backslash\!/^{j}
\end{equation}
and ${\bf J}$ is an arbitrary set of weights. In each css we consider the
system of {\em canonical} bases of ket vectors, $\{\kd{j,\lambda}\}\subset
{\cal K}^{j}$ and $\{\ku{j,\lambda}\}\subset \hat{\cal K}^{j}$,
$\lambda=-j,-j+1,...,j$, and the related bases of bra vectors satisfying
\begin{equation}
\bkud{j,\lambda}{j',\lambda'}=\delta^{j}_{j'}\delta^{\lambda}_{\lambda'}\,,
\quad
\bkdu{j,\lambda}{j',\lambda'}=\delta_{j}^{j'}\delta_{\lambda}^{\lambda'}\,.
\end{equation}
The metric operators of css,
\begin{equation}
\backslash\!/^{j}=\epsilon_{j}\sum_{\lambda}\ku{j,\lambda}\bu{j,\lambda}\,,
\end{equation}
are defined by the set of numbers $\epsilon_{j}=\pm 1$, $j\in {\bf J}$, that
gives the signature of the whole metric operator $\backslash\!/$ of the cvs
(\ref{(cvsrr)}). For each css the metric operator $\backslash\!/^{j}$ couples the
subspaces ${\cal K}^{j}=P^{j}{\cal K}$ and
$\hat{\cal K}^{j}={P^{j}}^{+}\hat{\cal K}$ given by the semi-hermitian
projection operator
\begin{equation}
P^{j}=\sum_{\lambda}\kd{j,\lambda}\bu{j,\lambda}.
\end{equation}
The spectral representations of the projections of the
operators $X\in su(2)$ are
\begin{equation}
X^{j}=P^{j}XP^{j}=\sum_{\lambda\lambda'}\kd{j,\lambda}X^{j}_{\lambda,
\lambda'}\bu{j,\lambda'},
\end{equation}
where $X^{j}_{\lambda,\lambda'}$ are the usual matrix elements of $X$ in the
canonical basis of the unitary irreducible representation of weight $j$.
Hereby it is easily to verify that the generators of the irreducible
representations with values in $L({\cal K}^{j}, {\cal K}^{j})$, denoted by
$/J^{j}_{a}/$, $a=1,2,3$, are semi-hermitian. Consequently, the generators
of the coupled representation are $\hat J_{a}^{j}=(J_{a}^{j})^{+}
\in L(\hat{\cal K}^{j}, \hat{\cal K}^{j})$.
\subsection{Finite-dimensional representations of $sl(2,\Comp)$}
The usual non-covariant generators of the $sl(2,\Comp)$ algebra are the
rotation generators, $I_{a},\, a=1,2,3$, and the Lorentz boosts,
$K_{a}$. From their well-known commutation rules it results that the algebras
generated by
\begin{equation}
M_{a}=\frac{1}{2}\left(I_{a}+iK_{a}\right)\,,\quad
N_{a}=\frac{1}{2}\left(I_{a}-iK_{a}\right)
\end{equation}
are two $su(2)$ algebras commuting with each other \cite{W}. Notice that
these algebras can not be seen as ideals of $sl(2,\Comp)$ since their
generators are complex linear combinations of the generators of a real
algebra.
Our aim is to construct the semi-unitary finite-dimensional representations
of $sl(2,\Comp)$ using our formalism. This means to consider from the
beginning that the generators of the coupled representations are
semi-hermitian satisfying $I_{a}^{+}=\backslash\!/ I_{a}/\!\backslash=\hat I_{a}$ and
$K_{a}^{+}=\backslash\!/ K_{a}/\!\backslash=\hat K_{a}$ (or $\overline{I_{a}}=I_{a}$ and
$\overline{K_{a}}=K_{a}$). Consequently, we must have
\begin{equation}\label{(mnv)}
M_{a}^{+}=\backslash\!/ N_{a}/\!\backslash=\hat N_{a}\,,\quad
N_{a}^{+}=\backslash\!/ M_{a}/\!\backslash=\hat M_{a}\,,\quad (\overline{M_{a}}=N_{a})\,.
\end{equation}
The general solution of this problem can be written starting with the space
of the reducible representation $(j_{1},j_{2})\oplus (j_{2},j_{1})$ of the
$sl(2,\Comp)$ algebra \cite{W}, ${\cal K}=({\cal K}^{j_{1}}\otimes
{\cal K}^{j_{2}})\oplus ({\cal K}^{j_{2}}\otimes {\cal K}^{j_{1}})$, with
$j_{1}\not=j_{2}$.
We find that the generators
\begin{eqnarray}
M_{a}&=&J_{a}^{j_{1}}\otimes P^{j_{2}} + J_{a}^{j_{2}}\otimes P^{j_{1}}\,,\\
N_{a}&=&P^{j_{1}}\otimes J_{a}^{j_{2}} + P^{j_{2}}\otimes J_{a}^{j_{1}}\,,
\end{eqnarray}
and the metric operator
\begin{eqnarray}
\backslash\!/=\epsilon_{j_{1},j_{2}}\sum_{\lambda\lambda'}\left(\ku{j_{1},\lambda}
\otimes \ku{j_{2},\lambda'}\bu{j_{2},\lambda'}\otimes \bu{j_{1},\lambda}
\right.\nonumber\\
\left.+ \ku{j_{2},\lambda'}\otimes
\ku{j_{1},\lambda}\bu{j_{1},\lambda}\otimes \bu{j_{2},\lambda'}\right)
\end{eqnarray}
with $\epsilon_{j_{1}j_{2}}=\pm 1$, satisfy equations (\ref{(mnv)}).
Hereby it results that the vector space coupled with ${\cal K}$ must be
$\hat {\cal K}=({\cal K}^{j_{2}}\otimes {\cal K}^{j_{1}})
\oplus ({\cal K}^{j_{1}}\otimes {\cal K}^{j_{2}})$.
These semi-unitary representations in cvs $({\cal K},\hat{\cal K},\backslash\!/)$
will be denoted by $[j_{1},j_{2}]$. They can be considered
{\em irreducible} in the sense of definition (\ref{(redu)}) since the metric
operator $\backslash\!/$ and the generators $M_{a}$ and $N_{a}$ do not have common
non-trivial invariant subspaces.
We can convince that with the help of the {\em chiral} projection operators,
$/P_{L}/= P^{j_{1}}\otimes P^{j_{2}}$ and $/P_{R}/=P^{j_{2}}\otimes P^{j_{1}}$,
which generalize the familiar left and right-handed ones of the theory of
Dirac spinors. These projection operators are just those of the invariant
subspaces of the generators $M_{a}$ and $N_{a}$. They form a complete set
of additive projection operators,
\begin{equation}
P_{L}P_{R}=0\,,\quad P_{L}+P_{R}=/\!\!/,
\end{equation}
but they are not semi-hermitian (or self-adjoint) operators since
\begin{equation}
{P_{L}}^{+}=\backslash\!/ P_{R}/\!\backslash\,,\quad
{P_{R}}^{+}=\backslash\!/ P_{L}/\!\backslash \,, \quad (\overline{P_{L}}=P_{R})\,.
\end{equation}
Therefore, the subspaces $P_{L}{\cal K}$ and $P_{R}{\cal K}$ are not
invariant subspaces of $\backslash\!/$ and the coupled representations $[j_{1},j_{2}]$
are irreducible from our point of view.
In the particular case of $j_{1}=j_{2}=j$ the solution is simpler.
The space ${\cal K}=\hat{\cal K}={\cal K}^{j}\otimes {\cal K}^{j}$ is just
that of the irreducible representation $(j,j)$ of $sl(2,\Comp)$ \cite{W}.
The generators have the form
\begin{eqnarray}
M_{a}&=&J_{a}^{j}\otimes P^{j}\\
N_{a}&=&P^{j}\otimes J_{a}^{j}
\end{eqnarray}
while the metric operator reads
\begin{equation}
\backslash\!/=\epsilon_{j}\sum_{\lambda\lambda'}\ku{j,\lambda}\otimes
\ku{j,\lambda'}\bu{j,\lambda'}\otimes \bu{j,\lambda}\,.
\end{equation}
These irreducible representations will be denoted by $[j]$.
\subsection{Rotation bases}
Now we can introduce the system of rotation dual bases in which the operators
$I^{2}=(I_{1})^{2}+(I_{2})^{2}+(I_{3})^{2}$ and $I_{3}$ as well as their
coupling partners are diagonal. Since $I_{a}=M_{a}+N_{a}$, the vectors of
these
bases can be constructed with the help of the Clebsh-Gordan coefficients of
the $SU(2)$ group \cite{AB}. The ket-down vectors of the rotation
basis of the subspace ${\cal K}^{j_{1}}\otimes {\cal K}^{j_{2}}\subset
{\cal K}$ are defined as
\begin{equation}
\kd{(j_{1},j_{2})\,s,\sigma}=\sum_{\lambda,\lambda'=\sigma-\lambda}\kd{j_{1},\lambda}\otimes
\kd{j_{2},\lambda'}\,\bk{j_{1},\lambda;j_{2},\lambda'}{s,\sigma}
\end{equation}
such that
\begin{eqnarray}
I^{2}\kd{(j_{1},j_{2})\,s,\sigma}&=&s(s+1)\kd{(j_{1},j_{2})\,s,\sigma}\,,\\
I_{3}\kd{(j_{1},j_{2})\,s,\sigma}&=&\sigma\kd{(j_{1},j_{2})\,s,\sigma}\,.
\end{eqnarray}
The rotation bases of the other spaces of our cvs have to be introduced in
the same manner. Then by taking into account that \cite{AB}
\begin{equation}
\bk{j_{1},\lambda;j_{2},\lambda'}{s,\sigma}=
(-1)^{s-j_{1}-j_{2}}\bk{j_{2},\lambda';j_{1},\lambda}{s,\sigma}
\end{equation}
and using the orthogonality relations of these coefficients, we
find the final forms of the metric operators in rotation bases. When
$j_{1}\not=j_{2}$ this is
\begin{eqnarray}
\backslash\!/=\epsilon_{j_{1},j_{2}}(-1)^{-j_{1}-j_{2}}\sum_{s=|j_{1}-j_{2}|}
^{j_{1}+j_{2}}(-1)^{s}\sum_{\sigma=-s}^{s}(
\ku{(j_{1},j_{2})\,s,\sigma}\bu{(j_{2},j_{1})\,s,\sigma}\nonumber\\
+\ku{(j_{2},j_{1})\,s,\sigma}\bu{(j_{1},j_{2})\,s,\sigma})\,,
\end{eqnarray}
while for $j_{1}=j_{2}=j$ we have
\begin{equation}\label{(omj)}
\backslash\!/=\epsilon_{j}(-1)^{2j}\sum_{s=0}^{2j}(-1)^{s}\sum_{\sigma=-s}^{s}
\ku{(j,j)\,s,\sigma} \bu{(j,j)\,s,\sigma}\,.
\end{equation}
In our opinion a good choice of the factors $\epsilon$ could be
\begin{equation}\label{(eee)}
\epsilon_{j_{1},j_{2}}=(-1)^{j_{1}+j_{2}-|j_{1}-j_{2}|}\,,\quad
\epsilon_{j}=(-1)^{2j}\,.
\end{equation}
Thus we obtain the spectral representations of the metric operators in
rotation dual bases. We observe that these bases are orthonormal
only for $j_{1}=j_{2}$. In the general case of $j_{1}\not=j_{2}$ the
ket-down vectors of the orthonormal basis of ${\cal K}$ are given by
the linear combinations
\begin{equation}
\kd{(\pm)\,s,\sigma}=\frac{1}{\sqrt{2}}(\kd{(j_{1},j_{2})\,s,\sigma}\pm
\kd{(j_{2},j_{1})\,s,\sigma})\,.
\end{equation}
Similarly we get the ket or bra vectors of the other orthonormal bases where
the metric operator can be represented as
\begin{eqnarray}
\backslash\!/=\epsilon_{j_{1},j_{2}}(-1)^{-j_{1}-j_{2}}\sum_{s=|j_{1}-j_{2}|}
^{j_{1}+j_{2}}(-1)^{s}\sum_{\sigma=-s}^{s}(
\ku{(+)\,s,\sigma}\bu{(+)\,s,\sigma}\nonumber\\
-\ku{(-)\,s,\sigma}\bu{(-)\,s,\sigma})\,.
\end{eqnarray}
From this formula we see that for $j_{1}\not=j_{2}$ the metric operator has
the symmetric signature $(n,n)$ with $n=(2j_{1}+1)(2j_{2}+1)$. The signatures
of the metric operators of the representations $[j]$ result from (\ref{(omj)})
and (\ref{(eee)}) to be either $(n,m)$ for integer $j$ or $(m,n)$ if $j$ is a
half-integer, where $n=(j+1)(2j+1)$ and $m=j(2j+1)$.
Hence for each pair of coupled irreducible representations of $sl(2,\Comp)$
the metric operator has a well-determined signature, $(n_{+},n_{-})$, which
shows us that the corresponding maximal symmetry group is just $U(n_{+},n_{-})$.
This result can be important from the physical point of view since the
transformations of this group are semi-unitary, leaving invariant the scalar
product of the carrier spaces of the $sl(2,\Comp)$ representations.
For example, in the theory of Dirac spinors the metric operator of the
representation $[1/2,0]$ has the signature $(2,2)$ which explains why the
whole algebra of $\gamma$ matrices and $sl(2,\Comp)$ generators is just
the $u(2,2)$ algebra \cite{TY}. We note that the corresponding $U(2,2)$ group
was recently considered as an extended gauge group of the theory of the Dirac
field in curved spacetime, obtaining thus new interesting results \cite{F}.
\section{Concluding remarks}
The presented formalism is the natural generalization of the Dirac's
bra-ket calculus to spaces with indefinite metric. The main point of our
proposal is to organize the four different spaces of the theory of general
tensors into a pair of coupled ket spaces mutually related with a pair of
coupled bra one. Then the metric operator can be correctly introduced in
accordance with the natural duality such that compatible scalar products and
dual forms do co-exist. In this way we recover the results of the theory of
the general linear transformations formulated simply in terms of operators,
independent on the concrete representations given by systems of dual bases.
For this reason, our approach helps one to precise the nature of the
different mathematical objects, avoiding the risk to confuse
among themselves those which accidently could have the same components or
matrix elements in several representations (e.g. the components of second
rank tensors and the operator matrix elements). Moreover, in this framework
some special operations used up to now only in particular problems get a
general meaning. We refer especially to the Dirac adjoint which in our
formalism can be defined for all the types of involved operators, taking
over the role of the Hermitian adjoint from the usual unitary case.
The presented example points out few of these advantages. Using our
general bra-ket calculus we are able to write the spectral representations
in different basis of the $sl(2,\Comp)$ generators and the metric operators of
the finite-dimensional irreducible representations with invariant scalar
products. The study of basis transformations, the generalization of the
chiral (left and right-handed) projection operators and the analyze of the
gauge group of the metric operator can be easily done in this context.
We tried here to remain in the spirit of the original
Dirac's bra-ket calculus, imagining a formalism with simple calculation
rules but having the ``memory" of the definitions and properties of its basic
elements. We hope that this should be appropriate for different applications
including algebraic programming on computers.
\subsection*{Acknowledgments}
I would like to thank Mircea Bundaru and Lucian Gligor for useful discussions
concerning the construction of the proposed formalism.
|
\section{Introduction }
It is well established that energy is supplied more or less continuously
to the outer lobes by beams of plasma ejected from the nucleus
at relativistic speeds. The radio jets, which are
the signatures of these energy-carrying beams,
expand rapidly with distance from the nucleus for the low-luminosity
FRI sources (Fanaroff \& Riley 1974), but are highly collimated and less
dissipative in the high-luminosity FRII objects (cf. Bridle \& Perley 1984).
However, jets in both these classes show evidence of flaring and recollimation
on different scales (cf. Hardee, Bridle \& Zensus 1996). The transverse widths
of the knots, $\Phi$, tend to increase with distance, $\Theta$, from the central or
nuclear component. However, the increase is not linear as would be expected in
a freely expanding jet. The spreading rate, defined to be the ratio $\Phi/\Theta$,
is usually larger close to the nucleus, which is followed by a slower spreading
rate or recollimation.
Some examples of well-studied jets in the high-luminosity FRII sources
are the ones in the 3CR quasars studied by Bridle et al. (1994, hereinafter
referred to as B94), the N
galaxy 3C390.3 by Leahy \& Perley (1995), the radio galaxies Cygnus A (cf. Carilli
et al. 1996) and 3C353 (Swain, Bridle \& Baum 1996). The transverse
knot widths in the inner jet in the quasar 3C351 suggest a high spreading
rate of 0.13, while the outer jet is better collimated. The jets
in the quasars 3C204, 3C263 and 3C334 exhibit faster than average spreading rate
when they are closest to the nucleus. The knots in the jets in 3C175 and 3C334
exhibit a decrease in their transverse widths before entering the hotspots, while
the jets in the quasars 3C204 and 3C263 show evidence of flaring before entering
the hotspots (B94).
In the N galaxy 3C390.3, the width of the jet is almost constant from 30 to 120 kpc,
it decreases till about 180 kpc and then increases again before
entering the hotspot (Leahy and Perley 1995). The width of the extended jet
in the radio galaxy 3C353 remains roughly constant beyond about 20 kpc
(Swain et al. 1996). Evidence of recollimation are also
seen in the small-scale nuclear jets. For example, in the
compact steep-spectrum quasar 3C138, Fanti et al. (1989) have reported evidence of
recollimation of the jet on scales of less than few hundred pc. The nuclear jet in
Cygnus A has a width of about 2.2 mas over a distance ranging from
2 to 20 mas from the nucleus (Carilli et al. 1996). These observations suggest
that radio jets
are often confined on scales ranging from the nuclear jets on the scale of parsecs
to the extended ones on scales of hundreds of kpc. The jets could be pressure
confined by the external environment or by magnetic fields around the jets
(cf. Sanders 1983; Bridle \& Perley 1984; Appl \& Camenzind 1992, 1993a,b;
Appl 1996; Kaiser \& Alexander 1997).
In this paper, we concentrate on the hotspots and radio knots in the jets and
use their sizes over a large
range of source sizes to investigate the collimation of jets in the compact
steep spectrum and larger sources. The hotspots indicate the
Mach disks where the jets terminate and subtend small angles
in the radio cores suggesting the high degree of collimation of the jets in
these sources (Bridle \& Perley 1984; B94; Fernini, Burns \& Perley 1997,
hereinafter referred to as F97).
Inclusion of hotspots enables us to study the collimation of jets for a large sample of sources
independent of whether a jet has been detected and its transverse width determined.
However, the jet momentum may be spread over a larger area than the cross-section of
the jet itself due to the dentist-drill effect discussed by Scheuer (1982),
where the end of the jet wanders about the leading contact surface drilling
into the external medium at slightly different places at different times.
Recent simulations of 3D supersonic jets suggest that cocoon turbulence
drives the dentist-drill effect (Norman 1996).
Over the last few years the sizes of hotspots have been determined for
samples of compact steep-spectrum radio sources using largely VLBI and MERLIN
observations, as well as for the larger objects using the VLA. Although
there is no well-accepted definition of a hotspot (cf. Laing 1989; Perley 1989),
we are interested in features which mark the collimation of jets
and have used the following empirical definition. The hotspots
are defined to be the brightest features in the lobes located further from
the nucleus than the end of any jet, and in the
presence of more extended diffuse emission these should be brighter by atleast a factor
of about 4 (cf. B94). In the presence of multiple hotspots, only the
primary hotspot has been considered. The jets which we consider have been well-mapped
and follow the defining criteria suggested by Bridle \& Perley (1984).
We have used the information on the sizes of the hotspots and the widths of the knots
in the jets to study the collimation and expansion of the radio jets over a
wide range of angular and linear scales.
The overall linear sizes of our objects are spread over about 5 orders of magnitude
ranging from about 50 pc to nearly 1 Mpc.
\input{jet_collimation_table.tex}
\section{Sample of sources}
Our sample has been chosen from well-defined samples of compact and
larger sources which have been observed with high angular resolution.
We have compiled our sample of compact sources from
the following samples of CSS and Gigahertz Peaked Spectrum or
GPS sources: (i) the sample of 67 sources listed by O'Dea (1998) and O'Dea \& Baum (1997)
which is based on the
Fanti et al. (1990) CSSs and Stanghellini et al. (1990, 1996) GPS objects;
(ii) the Sanghera et al. (1995) compilation of 62 objects, which consists of all sources
from the complete samples of 3C (Fanti et al. 1990), PW (Peacock \& Wall 1982)
and a Jodrell-Bank sample; and (iii)
the 7 confirmed compact symmetric objects or CSOs listed by Taylor,
Vermeulen \& Pearson (1995) and Taylor, Readhead \& Pearson (1996).
Our resulting combined sample consists of 86
objects, because there are many sources which are common to the
above lists, and have been counted only once in our combined list.
Of these 86, we have considered those which have a well-defined double-lobed
structure and where the hotspot sizes are available or could be determined. Sources with
a complex and highly distorted structure have been excluded. Typical examples of such sources are
3C48 (Wilkinson et al. 1991) and 3C119 (Ren-dong et al. 1991a). In addition, we have
excluded the source 0319+121 which has been listed as a GPS object by Stanghellini
et al. (1990), but single-epoch observations of the nucleus from 1.4 to 15 GHz show
that it has a flat radio spectrum (Saikia et al. 1998).
A few of our sources, such as 0108+388 (Baum et al. 1990),
have extended emission in addition to the compact double or triple structure
closer to the nucleus. The extended emission must be due to
earlier periods of activity, with the luminosity and size of the component being
governed by ageing and expansion of the component. On the other hand,
the compact structure closer to the nucleus represent more recent activity with
the size of the hotspots indicating the degree of collimation of the jets. Hence,
in such cases we have considered only the compact double or triple structure.
The sample which consists of 58 objects at this stage
has been further restricted to those which have been observed with at least 10
resolution elements, n$_b$, along the main axis. We also focus on the subsample which
has been observed with n$_b \geq$ 20, for a more reliable estimation of the
hotspot parameters. In our entire sample only 6 lobes have no hotspots which meet
our defining criteria. Excluding these lobes does not affect any of the conclusions
presented in this paper.
Sources which were earlier classified as CSS objects but were later found to be
of larger dimensions have not been excluded from the sample.
Our final sample of largely CSS and GPS objects consists of 39 sources, 9 of which
are $>$20 kpc. This sample is listed in
Table 1, which is arranged as follows: columns 1 and 2: source name and an
alternative name; column 3: optical identification where G denotes a galaxy,
Q a quasar and EF an empty field;
column 4: the sample where O, S and T denote O'Dea \& Baum (1997) and O'Dea (1998),
Sanghera et al. (1995) and Taylor et al. (1995, 1996) respectively;
column 5: redshift; column 6: the structural classification of the source where D
denotes a double, T a triple with a radio core, and T? a triple with a possible
core component; column 7: the angular separation of the hotspots on
opposite sides of the nucleus, expressed in arcsec; column 8:
the corresponding linear size in kpc
in an Einstein-de Sitter Universe with H$_o$=50 km s$^{-1}$ Mpc$^{-1}$;
column 9: the number of resolution elements, n$_b$, along the longest axis of the
source, which is the largest angular size of the radio source divided by the size of
the restoring beam along the axis of the source;
columns 10 and 11: the major and minor axes of one hotspot in mas;
columns 12 and 13: the
major and minor axes for the other hotspot in mas; column 14 : ratio of the separation
of the farther component from the radio core to the nearer one for those classified as
T or T?; column 15: references for radio structure. For
sources with a radio core, the size of the hotspot farther from the nucleus
is listed in columns 10 and 11. The hotspot sizes refer to the full width at half
maximum. For some of the images where the authors quote the size of the
entire component from the lowest reliable contour, we have estimated the sizes of
the hotspots from the images. These have been marked with an asterisk in the Table.
For comparison with larger sources we
consider those which have been observed with a similar number of resolution
elements, n$_b$, along the main axis of the source.
The comparison sample has
been compiled from the complete sample of 3CR sources (Laing, Riley \&
Longair 1983) and consists of the FRII sources which have been observed with
n$_b$ between about 10 and 40. These limits were chosen
so that the distribution of n$_b$ is similar to the CSS and GPS objects.
For the 3CR sources observed with the Cambridge 5-km telescope
we have confined ourselves to sources above a declination of 30$^\circ$ so
that the beams are not very elliptical. This sample is listed in Table 2
which is arranged similarly to Table 1 except for the following differences.
These are all triples from the 3CR sample, and hence the sample column and
structual information have been omitted.
In addition, we discuss briefly the collimation of jets using sources
which have been observed with a larger number of resolution elements along their axes.
These have been been observed with high resolution
and sensitivity with the Very Large Array (VLA) by B94,
Fernini et al. (1993), F97, Black et al. (1992), Leahy et al. (1997),
and Hardcastle et al. (1997, 1998). These observations give us estimates
of the sizes of hotspots in larger sources. B94 have reported observations of 13
quasars while F97 have listed hotspot sizes for 9 galaxies. Hardcastle et al.
(1998, hereinafter referred to as H98) have summarized the properties of jets,
cores and hotspots in the sample
of FRII galaxies with redshift $<$0.3 which have been observed by the Cambridge group.
In addition to the hotspots we have also considered the sizes of the knots in the jets listed in
Table 9 of B94 and of the knots in the CSOs observed by Taylor et al. (1995, 1996).
\begin{figure*}
\vspace{-5.8in}
\vbox{
\hbox{
\hspace{-0.5in}
\psfig{figure=fig1a.ps,width=7.1in}
\hspace{-3.6in}
\psfig{figure=fig1b.ps,width=7.1in}
}
\vspace{-6.0in}
\hbox{
\hspace{-0.5in}
\psfig{figure=fig1c.ps,width=7.1in}
\hspace{-3.6in}
\psfig{figure=fig1d.ps,width=7.1in}
}
}
\vspace{-0.2in}
\caption{The size of each hotspot is plotted against the largest angular size of
each source defined to be the separation of the oppositely-directed hotspots.
Filled circles indicate quasars while open circles
denote radio galaxies from the sample of largely CSS and GPS objects (Table 1).
The knots in the jets of the compact symmetric objects are denoted by
open squares. The hotspot sizes of the 3CR sources described in the text (Table 2) are indicated by
open and filled triangles for the galaxies and quasars respectively.
The arrows indicate upper limits.
The upper panels indicate all sources observed with at least 10 resolution elements
along the main axes while the lower panels show those which have been observed with at least
20 resolution elements along the main axes.
The solid lines denote the linear least-squares fit to the average hotspot size
for CSS and GPS sources, defined to
be those below 20 kpc. These linear fits have been extended beyond 20 kpc by a dotted line
to highlight the flattening of the relationship for larger sources.
The dashed lines denote the parabolic fit to the average hotspot size of the
sources in both the samples. For sources
with upperlimits, the hotspot sizes have been estimated by assuming the true values to
be close to the limits.
}
\end{figure*}
\section{Collimation of radio jets}
In the upper-left panel of Figure 1, we plot the angular sizes of the major and minor
axes of each hotspot against the
largest angular separation of each source in Tables 1 and 2, while
on the upper-right panel we present the equivalent linear sizes for
the same sources. The lower panels show the same plots but only
for those sources which have been observed with at least about 20 resolution elements along
the long axis of the source.
Since we sometimes do not have information on the radio core, especially for the
CSS and GPS objects we have plotted the hotspot sizes against the overall
separation of the oppositely-directed hotspots.
In this figure, we have also plotted the knots in the CSOs imaged by Taylor
et al. (1995, 1996).
In this paper, we have made the least-squares fits to the hotspots
by taking the average hotspot size for each source. This is defined to be the
geometric mean of each hotspot and, for those with hotspots on both sides,
the average of the two oppositely-directed hotspots for each source.
As mentioned earlier, a distance of 20 kpc was chosen as the canonical
limit for the size of the CSS and GPS objects (cf. Fanti et al. 1990; O'Dea 1998).
A linear least-squares fit to the hotspots for the CSS and GPS objects in Figure 1
observed with at least 10 resolution elements shows that if the jet width
is expressed as $d_{jet} = k l^{n}$, then $n \sim 1.02\pm0.06$ and log$(k) \sim -1.23\pm0.05$.
For those observed with at least 20 resolution elements the value of
$n$ is $1.10\pm0.07$ and log$(k) \sim -1.4\pm0.05$. The sizes of the knots in the CSOs
are consistent with this trend.
There appears to be a well-defined relationship with
the hotspot size increasing with the total size of the source upto a distance of about
20 kpc, the canonical limit for the sizes of CSSs. The sizes of the knots in the jets
are similar to those of the hotspots. This relationship is consistent with
self-similar models for the evolution of radio sources during the CSS and GPS phase
(cf. Kaiser \& Alexander 1997).
\subsection{Comparison with larger 3CR sources observed with
similar resolution elements}
To investigate this relationship for sources $>$ 20 kpc,
we have compared the CSS and GPS sources (Table 1)
with the sample of 3CR sources (Table 2) which have been observed
with similar resolution relative to the largest angular size as used for the
CSS sources. The number of resolution elements, n$_b$, for the CSS
sources vary from about 10 to 77 with a median value of about 19 while for
the 3CR sources
n$_b$ ranges from about 10 to 39 with a median value of about 23. Only 5 of the
CSSs have n$_b$ $>$40. Considering the sources which have been observed with
n$_b$ between about 20 and 40, the median values of n$_b$ for CSS and GPS, and 3CR
sources are 25 and 28 respectively, and a Kolmogorov-Smirnov test shows that
the two distributions are the same with a probability of 66 per cent.
The redshifts of the CSS and GPS sources range from about
0.1 to 2 with a median value of about 0.6 while the redshifts for the 3CR
sample is somewhat larger ranging from about 0.16 to 2.0, with a median
value of about 0.8.
The extension of the linear least-squares fit to the
CSS sources beyond 20 kpc shows that most of the larger sources
observed with a similar value of n$_b$ lie below this line,
suggesting a flattening of this relationship. This indicates recollimation
of the jets beyond $\sim$20 kpc.
Since most of the measured sizes are upper limits, it is difficult
to determine the precise nature of the relationship from the present
sample. Assuming that the sizes of the hotspots with upper limits are
close to these limits, we have attempted a linear least-squares as well
as a parabolic fit to the entire data. The parabolic fits are significantly
better, suggesting again a flattening of the relationship. To confirm the
trend and determine the degree of flattening, we
need to measure more precisely the sizes of the hotspots, rather than have limits
to their sizes.
\subsection{3CR sources observed with larger values of n$_b$}
Over the last few years several authors have determined the sizes of the hotspots
of 3CR FRII sources from high-resolution VLA observations
with n$_b$, the number of resolution elements along the axis of the source,
in the range of about 35 to 1040 (e.g. B94; F97; Black et al. 1992;
Leahy et al. 1997; Hardcastle et al. 1997). With higher values of n$_b$ one should
expect to see smaller-scale structures within the hotspots, and hence a decrease in
the size of the hotspots. A linear least-squares fit to
the hotspot size $-$ n$_b$ diagram for the B94, F97 and H98 samples
exhibits a weak trend for the hotspot size to decrease with n$_b$ with a
slope of $-0.15\pm0.10$. Because of different definitions of hotspots and also
difficulties in measuring their sizes, detailed comparisons are often difficult and
contentious. However, the sizes of the hotspots from the B94, F97 and H98 have
median values in the range of about 1.9 to 4.5 kpc and are consistent with a flattening of
the hotspot size $-$ linear size relationship beyond about 20 kpc. The knots in the
jets (Table 9 of B94) are also consistent with this flattening.
\begin{figure}
\vspace{-5.8in}
\hbox{
\hspace{-0.5in}
\psfig{figure=fig2.ps,width=7.1in}
}
\vspace{-0.2in}
\caption{The ratio, r$_{hs}$, of the size of the hotspot farther from the radio core to
that of the oppositely-directed one closer to the core, is plotted against the
separation ratio, r$_d$, defined to be the ratio of the separations of the corresponding
hotspots from the core. The open and filled circles denote galaxies and quasars from
the sample of largely CSS and GPS objects (Table 1), the open and filled stars denote
galaxies and quasars from B94 and F97, while the triangles represent radio galaxies from
H98. Filled and open squares indicate quasars and radio galaxies from Table 2 which
have not been already plotted.
}
\end{figure}
\begin{figure}
\vspace{-5.5in}
\hbox{
\hspace{-0.5in}
\psfig{figure=fig3.ps,width=7.1in}
}
\vspace{-0.2in}
\caption{The ratio, r$_{hs}$, of the size of the hotspot farther from the radio core to
that of the oppositely-directed one closer to the core, is plotted against the
total luminosity at 5 GHz. The symbols are as defined in Figure 2.
}
\end{figure}
\subsection{Opening angle of the jets}
Using the geometric mean of the major and minor axes for each hotspot and the average
of the hotspots for each source, we estimate the mean opening angle, ($\phi_{obs}$),
for sources below 20 kpc to be about 7$^\circ$. For sources observed with
atleast 20 beam widths along the source axes the mean opening angle is about 5$^\circ$.
The estimates of the mean opening angle is roughly consistent with estimates for
individual jets in different angular and linear scales. It is relevant to note
that the opening angle
estimated from the hotspots could, in principle, be larger than those estimated
from knots in the jets because of the dentist drill effect (Scheuer 1982; Norman 1996). In
the radio galaxy Cygnus A, the opening angle of the jet determined from the knots
as well as the low surface brightness inter-knot regions is $\approx$1.6$^\circ$
(Carilli et al. 1996), although the morphology of the jet is consistent with
regions of larger opening angle of $\approx$5$^\circ$ (Perley, Dreher \& Cowan 1984).
The jet in M87 is well-collimated on all scales out to about 2 kpc. The
innermost VLBI structure indicates an opening angle of 18$^\circ$, while VLA
observations indicate that between the jet and knot A, the jet appears as
a uniformly expanding cone with an opening angle of 7$^\circ$ (Biretta 1996;
Biretta \& Junor 1995). Among the core-dominated radio sources, one of the
best studied ones is the quasar 3C345 which has an opening angle of 26$^\circ$
(Zensus, Cohen \& Unwin 1995). Other examples are NRAO 140 with an opening
angle of 12$^\circ$ (Marscher 1988; Unwin \& Wehrle 1992) and 0836+710
which has a mean opening angle of 4.4$^\circ$ (Hummel et al. 1992).
\subsection{Asymmetry in collimation}
We have also examined the dependence of the ratio of the sizes of the two
oppositely-directed hotspots, r$_{hs}$, on their relative separations from the nucleus.
In Figure 2 we plot the ratio, r$_{hs}$,
defined to be the ratio of the size of the farther hotspot to that of the nearer one, against
the separation ratio, r$_d$, for all the sources with
a detected radio core and hotspots on both sides. The separation ratio, r$_d$, is
defined to be $\geq$ 1, while
r$_{hs}$ is the ratio of the geometric mean of the corresponding hotspots. We have
used only those hotspots without an upperlimit to any of their axes and with a detected
radio core from the sample of largely CSS and GPS objects (Table 1), the 3CR objects
listed in Table 2 and the observations of B94, F97 and H98.
For the 33 high-luminosity sources with a radio luminosity at 5 GHz
$\geq 10^{26}$ W Hz$^{-1}$ sr$^{-1}$, 24 have r$_{hs}$ $>$ 1, implying that
the farther hotspot is larger. This trend is
also seen in the most asymmetric objects, defined to be those with r$_d >$ 2.
Of the 12 sources with r$_d >$ 2, 9 have r$_{hs}$ $>$ 1, and almost all these objects are CSSs.
Either light-travel time effects or an intrinsic or environmental asymmetry can
produce the effect in the observed sense.
For a source inclined at about 50$^\circ$ to the
line-of-sight and a hotspot speed of about 0.5c, r$_d$ is close to about 2.
There have been several estimates of hotspot advance speeds in VLBI-scale double sources.
O'Dea (1998) lists 7 objects which have upper limits to component proper motion
which are subluminal, ranging from 0.05-0.5c. Since O'Dea's review the upper limit
for 1934$-$638 has been revised to $\sim$0.03$\pm$0.2c (Tzioumis et al. 1998), and
an upper limit for 1607+26 (CTD93) has been reported to be less than 0.35c (Shaffer \&
Kellermann 1998).
Current estimates of hotspot speeds suggest that the effects of an asymmetry dominate.
Almost all these objects with r$_d >$2 are compact steep spectrum radio sources, and the
deficit of objects with r$_d >$2 and r$_{hs} <1$ is possibly a reflection of their
evolution in an asymmetric environment.
The tendency for r$_{hs}$ $>$ 1 is not seen in objects of lower luminosity (Figure 3). For example,
in the 50 sources with radio luminosity at 5 GHz less than 10$^{26}$ W Hz$^{-1}$ sr$^{-1}$,
only 24 have r$_{hs}$ $>$ 1. A number of these objects have r$_{hs}$ significantly smaller
than 1, which could be due to an intrinsic asymmetry in the collimation of jets on
opposite sides of the nucleus.
\section{Discussion and concluding remarks}
We summarise the principal trends reported in this paper.
\begin{enumerate}
\item The relationship between the hotspot sizes and the overall size of
the CSS and GPS sources and studies of individual knots in the jets in
these sources, suggest that they evolve in an approximately self-similar way.
The hotspot size increases with distance from the core as $d_{jet}
\propto l^{n}$ where
$n \sim$ 1.0 for sources smaller than about 20 kpc. This is similar for sources
which have been observed with at least either 10 or 20 resolution elements, n$_b$, along
the main axes so that the hotspots could be identified reliably.
\item For larger sources observed with at least either 10 or 20 resolution elements along the
main axes and with a similar distribution of n$_b$ to the sample of largely
CSS and GPS sources, there appears to be a
flattening of the relationship. However, since most of the sources have upper limits
to their hotspot sizes the precise slope could not be determined reliably.
\item For samples of sources observed by Bridle et al. (1994) and Fernini et al. (1993, 1997)
with n$_b$ in the range of about 40 to 200, and a sample of 3CR sources with z$<$0.3 summarized by
Hardcastle et al. (1998) with n$_b$ in the range of about 35 to 1040, the dependence of
hotspot size on n$_b$ is weak. The hotspot size varies by a factor of about 2 for an
increase in n$_b$ by a factor of about 300. The hotspot sizes of these large sources
are consistent with a flattening of the hotspot size-linear size relationship
seen for CSS and GPS objects.
\item There is a tendency for the farther hotspot to be larger in sources with a luminosity at
5 GHz $>$ 10$^{26}$ W Hz$^{-1}$ sr$^{-1}$.
This could be caused by both light travel time effects as well as an asymmetric environment.
No such trend is seen in objects of lower luminosity where the nearer hotspot is significantly larger in
a number of sources. This could be due to an asymmetry in collimation on opposite sides of the nucleus.
\item There is a trend for the farther hotspot to be larger
in the most asymmetric objects, defined to be those with r$_d >$ 2.
Almost all the objects with r$_d >$ 2 are CSS objects. Current estimates of hotspot advance
speeds suggest that these trends are due to an intrinsic asymmetry in the environment
rather than light travel time effects.
\end{enumerate}
In the CSS phase, the variation of the sizes of the hotspots and
the widths of the knots in the jets with linear size
could be due to the ambient pressure falling with distance from the nucleus,
if the jet is pressure confined.
The confinement of the jet by the ambient medium depends on
whether the jet pressure is comparable to the external pressure.
The equipartition pressure in the hotspots or knots varies from about 10$^{-6}$ N m$^{-2}$
on scales of a few tens of parsec to about
$10^{-9}$ N m$^{-2}$ at about 10 kpc from the nucleus (Fanti et al. 1995; Readhead 1995;
Readhead et al. 1996a,b). For large sources, the equipartition pressure in the knots of the jets
is in the range of about $10^{-10}$ to $10^{-12}$ N m$^{-2}$ (Potash and Wardle 1980; Bridle et al. 1994).
Although the nature of the confining medium within about 20 kpc, i.e. the CSS phase,
needs to be better understood (cf. Fanti et al. 1995), the gaseous components seen in
absorption at X-ray, UV and optical wavelengths (Mathur et al. 1994; Elvis et al. 1996;
Netzer 1996) could play an important role in addition to the broad-line and narrow-line gas.
For the larger FRII sources the flatter relationship
seen in the present data suggests recollimation of the jets beyond the CSS phase.
It is important to understand the physical process responsible for the
recollimation of jets. Two interesting scenarios
which have suggested for recollimation are
hydrodynamic collimation involving shocks (Sanders 1983; Falle \& Wilson 1985;
Komissarov \& Falle 1996) and magnetic collimation (Begelman, Blandford \& Rees 1984;
Appl \& Camenzind 1992, 1993a,b; Appl 1996). The latter could provide a natural explanation
of a constant width in large sources, which appears to be suggested by some of the
observations such as those of B94 and F97, and in 3C353 (cf. Swain et al. 1996).
\section*{Acknowledgments}
We thank Robert Laing, Peter Scheuer and an anonymous referee for their critical comments and
suggestions, and our colleagues at NCRA
especially Gopal-Krishna and late Vijay Kapahi for
their comments on this piece of work, and Judith Irwin, Ishwara-Chandra and
Vasant Kulkarni for their detailed comments on the manuscript. This research has made use of the
NASA/IPAC extragalactic database (NED)
which is operated by the Jet Propulsion Laboratory, Caltech, under contract
with the National Aeronautics and Space Administration.
|
\section{Introduction}
Any dynamical evolution that occurs in nature is affected by noise.
In a neuronal system the noise might be comparable in magnitude to
purported underlying deterministic dynamics; in celestial
mechanics the degrees of freedom omitted from a particular
set of equations may be accounted for by very weak noise.
Our task here and in two preceding papers\rf{noisy_Fred,conjug_Fred}
is to systematically account for the effects of noise on measurable
properties such as dynamical averages\cite{bene} in classical chaotic dynamical
systems.
The theory is also closely related to the semiclassical expansions based
on Gutzwiller's formula for the trace in terms of classical periodic
orbits\rf{gutbook} in that both are perturbative theories (in the noise
strength or $\hbar$) derived from saddlepoint expansions of a path
integral containing a Cantor set of unstable stationary points (typically
periodic orbits). The analogy with quantum mechanics and
field theory has been made explicit in\rf{noisy_Fred} where
Feynman diagrams were used to find the lowest nontrivial noise corrections.
Unfortunately like its quantum counterpart, the Feynman diagram method
for stochastic dynamics quickly becomes unwieldy at higher orders; rather
than applying it directly we turn the argument around and suggest that
the more efficient recent approaches of\rf{conjug_Fred} and the present
paper be applied to difficult perturbative problems of quantum mechanics
and field theory.
An elegant method, inspired by the classical perturbation theory of
celestial mechanics, is that of smooth conjugations\rf{conjug_Fred}.
In this approach
the neighborhood of each saddlepoint is flattened by an appropriate
coordinate transformation, so the focus shifts from the original dynamics
to the properties of the transformations involved. An elementary example
is the Ulam map $f(x)=4x(1-x)$ which is solved exactly by the transformation
$x=\sin^2(\pi\theta/2)$ leading to the piecewise linear tent map
$f(\theta)=1-|1-2\theta|$. In general there is no such explicit solution,
but the expressions obtained for perturbative corrections
are much simpler than those found
from the equivalent Feynman diagrams. Using these techniques, we were
able to extend the stochastic perturbation theory to the fourth order in the
noise strength.
Fourth order should be sufficient for most realistic calculations, but
does not provide enough information to determine the convergence properties
of the expansion, or determine eigenvalues beyond the first few.
In this paper we develop
a third approach, based on construction of an explicit matrix
representation of the stochastic evolution operator.
Numerical implementation requires a
truncation to finite dimensional matrices,
and is less elegant than the smooth conjugation method,
but for high expansion orders (here eighth, but higher orders seem quite
feasible) and many eigenvalues it is currently unsurpassed.
As with the previous formulations, it retains the periodic orbit structure,
thus inheriting valuable information about the dynamics.
In the following sections we define the stochastic dynamics and show how
to obtain matrix representations, both globally and located on the
periodic orbits, as an expansion in terms of the noise strength $\sigma$.
The matrix elements are obtained from derivatives of the dynamics computed
around each periodic orbit. We give as a numerical example the quartic
map considered in both previous papers, although the approach is very
general and is by no means restricted to one dimension, to maps, or to
Gaussian noise. We find that up to eighth order, the cumulants converge
super-exponentially with the length of periodic orbit and the expansion is
now shown to be accurate to larger values of $\sigma$.
\section{The stochastic evolution operator and its spectrum}
An individual trajectory in presence of additive noise is generated
by iteration
\beq
x_{n+1}=f(x_{n})+\sigma\xi_{n}
\,,
\ee{mapf(x)-Diag}
where $f(x)$ is a map,
$\xi_n$ a random variable with the normalized
distribution $p(\xi)$,
and $\sigma$ parametrizes the noise strength.
In what follows we shall assume that the mapping $f(x)$ is
one-dimensional and expanding, and that the $\xi_n$ are uncorrelated.
A density of trajectories $\phi(x)$ evolves with time as
\beq
\phi_{n+1}(y) =
\left(
\Lop
\circ
\phi_{n}\right)(y)
= \int dx \, \Lop(y,x) \phi_{n}(x)
\ee{DensEvol}
where $\Lop$ is the {\evOper}
\bea
\Lop(y,x) &=& \delta_\sigma(y-f(x))
\continue
\delta_\sigma(x)
&=& \int \delta(x-\sigma \xi) p(\xi) d\xi
\,=\, \frac{1}{\sigma} p\left( \frac{x}{\sigma} \right)
\,.
\label{oper-Diag}
\eea
For a repeller the leading eigenvalue of the {\evOper}
yields a physically measurable property of the dynamical system,
the escape rate from the repeller.
In the case of deterministic flows, the periodic orbit theory
yields explicit formulas for the spectrum of $\Lop$ as zeros
of its {\fd}\rf{QCcourse}. Our goal here
is to explore the extent to which such methods are applicable to
systems with noise and to quantum systems. In particular, we are
interested in exploring the dependence of the eigenvalues
$\eigenvL(\sigma)$ of $\Lop$
on the noise strength parameter $\sigma$.
The eigenvalues are determined by the eigenvalue condition
\beq
\eigCond(\sigma,\eigenvL(\sigma)) = \det(1-\Lop/\eigenvL(\sigma)) =0
\label{eigCond}
\eeq
where
$
\eigCond(\sigma,1/z) = \det(1-z\Lop)
$
is the {\fd} of the {\evOper} $\Lop$.
Computation of such determinants
commences with evaluation of the traces of powers of
the {\evOper}
\bea
\tr {z\Lop \over 1-z\Lop}
&=&
\sum_{n=1}^{\infty} C_{n} z^n
\,,\qquad C_{n} = \tr \Lop^n
\,,
\label{tr-L-ith-Diag}
\eea
which are then used to compute the cumulants
$Q_{n}=Q_{n}(\Lop)$ in the cumulant expansion
\beq
\det(1-z\Lop)
= 1- \sum_{n=1}^{\infty} Q_{n} z^n
\,,
\ee{Fred-cyc-exp-Diag}
by means of the recursion formula
\beq
Q_n = {1 \over n}\left( C_n -C_{n-1} Q_1 - \cdots C_1 Q_{n-1}\right)
\ee{Fd-cyc-exp-Diag}
which follows from the relation
\beq
\det(1-z\Lop)
\defeq
\exp\left(-\sum_n^\infty {z^n \over n} \tr \Lop^n \right)
\,.
\ee{det-tr-Diag}
Our task is to compute the cumulants $Q_n$.
We start by introducing a matrix representation for $\Lop$.
\section{Matrix representation of evolution operator}
\label{ENTIRE}
As the mapping $f(x)$ is expanding by assumption, the {\evOper} \refeq{DensEvol} smoothes the initial
distribution $\phi(x)$. Hence it is natural to assume that
the distribution $\phi_{n}(x)$ is analytic, and represent it as
a Taylor series, intuition being that the action of $\Lop$ will
smooth out fine detail in initial distributions and the expansion
of $\phi_{n}(x)$ will be dominated by the leading terms in the series.
An analytic function $g(x)$ has a Taylor series expansion
\[
g(x) = \sum_{m=0}^\infty
\frac{x^{m}}{m!} \left. \frac{\pde^{m}}{\pde y^{m}}g(y) \right|_{y=0}
\,.
\]
Expanding $\Lop(y,x)$ in Taylor series in $y$
enables us to rewrite traces of $\Lop^n$ as
\bea
\tr \Lop^2
&=&
\int dx dy\,
\Lop(y,x)
\Lop(x,y)
\continue
&=&
\sum_{m,m'} \int dx dy\,
\left(\frac{y^{m'}}{m'!}
\left. \frac{\pde^{m'}}{\pde v^{m'}}
\Lop(v,x)
\right|_{v=0}\right)
\ceq \qquad\qquad
\left(\frac{x^{m}}{m!}
\left. \frac{\pde^{m}}{\pde u^{m}}
\Lop(u,y)
\right|_{u=0}\right)
\nnu
\eea
Following H.H.~Rugh\rf{Rugh92} we
now define the matrix ($m,m'= 0,1,2, ...$)
\beq
\left(\Lmat{}\right)_{m'm} =
\left. \frac{\pde^{m'}}{\pde y^{m'}}
\int dx \, \Lop(y,x)
\frac{x^{m}}{m!} \right|_{y=0} .
\ee{Lmat-Diag}
$ \Lmat{}$ is a matrix representation of $\Lop$ which
maps the $x^m$ component of the density of trajectories $\phi_n(x)$ in \refeq{DensEvol} to the $y^{m'}$ component of the density $\phi_{n+1}(y)$, with $y=f(x)$.
The desired traces can now be evaluated as traces of the matrix
representation $ \Lmat{}$,
$
\tr \Lop^n
=
\tr \Lmat{}^n
\,.
$
As $\Lmat{}$ is infinite dimensional, in actual
computations we have to truncate it to a given finite order.
The Feynman diagrammatic and the smooth conjugation methods developed
in the preceding papers\rf{noisy_Fred,conjug_Fred}
require no such approximations. However,
as we shall see below, for expanding flows the structure of $ \Lmat{}$
is such that its finite truncations give very accurate spectra.
Our next task is to evaluate the matrix elements of $ \Lmat{}$.
\section{Weak noise expansion of the evolution operator}
We have written the operator $\Lop$ in \refeq{oper-Diag} in terms of
the Dirac delta function,
$
\Lop(x',x)=\int \delta(x'-f(x)-\sigma\xi)p(\xi)d\xi
$,
in order to emphasize that
in the weak noise limit the stochastic trajectories are concentrated
along the deterministic trajectory $x' = f(x)$.
Hence it is natural to
expand the delta function in a Taylor series
in $\sigma$
\bea
\Lop(x',x) &=&
\delta(x'-f(x))
\ceq
+\,
\sum_{n=2}^{\infty}\frac{(-\sigma)^n}{n!}\delta^{(n)}(x'-f(x))
\int \xi^n p(\xi)d\xi
\,,
\nnu
\eea
where
$
\delta^{(n)}(y) = {\pde^n \over \pde y^n} \delta(y)
\,.
$
This yields a representation of the
{\evOper} centered along the deterministic trajectory, with
the
{\FPoper} $\delta(x'-f(x))$, and corrections given by derivatives
of delta functions weighted by moments of the noise
distribution $a_n=\int p(\xi)\xi^nd\xi$,
\begin{equation}
\Lop(x',x)
=
\delta(x'-f(x)) \,+\,
\sum_{n=2}^{\infty}\frac{(-\sigma)^n}{n!} a_n \delta^{(n)}(x'-f(x)).
\label{opexp}
\end{equation}
In our numerical tests we find it convenient to assume that the noise
is Gaussian,
$
p(\xi)=
e^{-\xi^2/2}/{\sqrt{2\pi}}
\,.
$
For the Gaussian noise all $a_n$ moments are known,
and the weak noise expansion of $\Lop$ is
\bea
\Lop(x',x)
&=&
{1 \over \sqrt{2 \pi \sigma^2}} e^{-{(x'-f(x))^2/2\sigma^2} }
\continue
&=&
\sum_{n=0}^{\infty}
\frac{\sigma^{2n}}{n!2^n} \delta^{(2n)}(x'-f(x))
\continue
&=&
\delta(x'-f(x)) + {\sigma^2 \over 2} \delta^{(2)}(x'-f(x))
\ceq \qquad\qquad
+ {\sigma^4 \over 8} \delta^{(4)}(x'-f(x)) + \cdots
\,.
\label{delGaussExp}
\eea
The choice of Gaussian noise is not essential,
as the methods that we develop here apply
equally well to any other peaked smooth noise distribution, as
well as space dependent noise distributions $p(x,\xi)$.
In any case, as the neighborhood of any trajectory is nonlinearly distorted by the flow, the integrated noise is never Gaussian, but colored.
\section{Local matrix representation of evolution operator}
\label{LocMatr-Diag}
Traces of powers of the {\evOper} $\Lop^n$ are now also a power
series in $\sigma$, with contributions composed of
$\delta^{(m)}(f(x_a)-x_{a+1})$ segments.
The contribution is non-vanishing only if the sequence
$x_1,x_2,...,x_n, x_{n+1} = x_{1}$ is a periodic orbit of
the deterministic map $f(x)$.
Thus the series expansion of $\tr \Lop^n$ has support on
all periodic points $x_a = x_{a+n}$ of period $n$,
$f^n(x_a)=x_a$; the skeleton of periodic points of the deterministic
problem also serves to describe the weakly stochastic flows.
The contribution of the
$x_a$ neighborhood
is best presented by introducing a coordinate system $\field_a$
centered on the cycle points,
together with a notation for the
map \refeq{mapf(x)-Diag} and the operator \refeq{oper-Diag}
centered on the
$a$-th cycle point
\bea
x_a &\to& x_a+\field_a \,, \qquad a=1,...,n_p
\continue
f_a(\field) &=& f(x_a+\field)
\continue
{\Lop}_a(\field_{a+1},\field_a)
&=&
\Lop(x_{a+1}+\field_{a+1},x_a+\field_a)
\,.
\eea
The weak noise expansion \refeq{opexp} for the $a$-th segment operator
is given by
\[
{\Lop}_a(\field',\field)=\sum_{n=0}^{\infty}\frac{(- \sigma)^{n}}{n!}
a_n \delta^{(n)}(\field'+x_{a+1}- f_a(\field))
\,.
\]
Repeating the steps that led to \refeq{Lmat-Diag}
we construct the local matrix representation of $\Lop_{a}$ centered on
the $x_a \to x_{a+1}$ segment of the deterministic trajectory
\bea
\left(\Lmat{a}\right)_{m'm}
&=&
\left. \frac{\pde^{m'}}{\pde \field'^{m'}}
\int d\field \,\Lop_{a}(\field',\field)
\frac{\field^{m}}{m!} \right|_{\field'=0} .
\continue
&=&
\sum_{n=max(m-m',0)}^{\infty}
\frac{(-\sigma)^n}{n!}a_n(\Bmat{a})_{m'+n,m}
\,.
\label{BtoL}
\eea
Due to its simple dependence on the Dirac delta function,
$\Bmat{}$ can expressed in terms of derivatives of the inverse
of $f_a(\field)$:
\bea
(\Bmat{a})_{n m}
&=&
\left. \frac{\pde^{n}}{\pde \field'^{n}}
\int d\field \,
\delta(\field' + x_{a+1} -f_a(\field))
\frac{\field^{m}}{m!}
\right|_{\field'=0}
\continue
&=&
\left. \frac{\pde^{n}}{\pde \field'^{n}} \frac{(f_a^{-1}(x_{a+1}+\field')-x_a)^{m}}{m!|f_a'(f_a^{-1}(x_{a+1}+\field'))|}\right|_{\field'=0}
\continue
&=&
\frac{\mbox{sign}(f_a')}{(m+1)!}\left.\frac{\pde^{n+1}(
{\cal F}_a(\field')^{m+1})}{\pde \field'^{n+1}}\right|_{\field'=0},
\label{Bmatrix}
\eea
where we introduced the shorthand notation ${\cal F}_a(\field')=
f_a^{-1}(x_{a+1}+\field')-x_a$.
If we expand ${\cal F}_a(\field')$ in a Taylor series, the constant term
is
zero, since $f_a^{-1}(x_{a+1})=x_a$. So we can write:
\bea
{\cal
F}_a(\field')=\sum_{l=1}^{\infty}\frac{{\cal
F}_a^{(l)}}{l!}\field'^l,
\eea
where $1/{\cal F}_a^{(1)}=f'_a$.
The matrix elements can be calculated explicitly as a multinomial expansion \cite{abramo}
\bea
\left(\sum_{l=1}^{\infty}\frac{x_l}{l!}t^l\right)^m & = & m!\sum_{n=l}^{\infty}
\frac{t^n}{n!} \continue
\ceq \cdot \, \sum(n|a_1,...,a_n)'x_1^{a_1}...x_n^{a_n},
\label{Abram}
\eea
where the second sum $\left(\sum\right)$ goes over all non-negative integers
such that:
\bea
a_1+2a_2+...+na_n=n, \mbox{\hspace{0.3cm}} a_1+a_2+...+a_n=m,
\eea
and the multinomial coefficient is:
\bea
(n|a_1,a_2,...,a_n)'=\frac{n!}{(1!)^{a_1}a_1!(2!)^{a_2}a_2!...(n!)^{a_n}a_n!}.
\eea
We apply the formula $\refeq{Abram}$ to ${\cal F}_a(\field')$ with power $m+1$:
\bea
({\cal F}_a(\field'))^{m+1} &=& (m+1)! \sum_{l=m+1}^{\infty}\frac{\field'^n}{n!}\sum(l|a_1,a_2,...,a_l)'
\continue
\ceq \mbox{} \cdot \, ({{\cal F}_a}^{(1)})^{a_1}({{\cal F}_a}^{(2)})^{a_2}...({{\cal F}_a}^{(l)})^{a_l}.
\eea
For the $(n+1)$ -th derivative of this expression evaluated at $\field'=0$ only the $l=n+1$ term is non-vanishing.
The matrix elements vanish for $n<m$, so $\Bmat{}$ is
a lower triangular matrix:
\bea (\Bmat{a})_{n m} &=& \sum(n+1|a_1,a_2,...,a_{n+1})'
\continue
\ceq \cdot \, ({{\cal F}_a}^{(1)})^{a_1}({\cal F}_a^{(2)})^{a_2}...({\cal F}_a^{(n+1)})^{a_{n+1}}.
\eea
The diagonal and the nearest off-diagonal matrix elements
can easily be worked out. Here we show the first four expressed in terms
of the derivatives of the original map:
\bea
(\Bmat{a})_{m m}~~~ &=& \frac{1}{|f_a'|f_a'^{m}}
\continue
(\Bmat{a})_{m+1,m} &=& - \frac{1}{2} \frac{(m+2)!}{m!} \frac{f_a''}{|f_a'|f_a'^{m+2}}
\label{Bexplicit}\\
(\Bmat{a})_{m+2,m} &=&
-\frac{(m+3)!}{24 m! |f_a'|f_a'^{m}}
\left( \frac{f_a'''}{f_a'^{3}} -3 (m+4) \frac{(f_a'')^2}{f_a'^{4}}
\right)
\continue
(\Bmat{a})_{m+3,m} &=&
- \frac{(m+4)!}{48 m!}{|f_a'|f_a'^{m}}
\left( 2 \frac{f_a''''}{f_a'^{4}} - 4(m+5) \frac{f_a'' f_a'''}{f_a'^{5}}
\right.
\ceq \qquad\qquad
\left.
+ (m+5)(m+6) \frac{f_a''^{3}}{f_a'^{6}}
\right)
\continue
& &\cdots
\nnu
\,,
\eea
where $f_a'$, $f_a''$,
$\cdots$ refer to the derivatives of $f(x)$ evaluated at the
periodic point $x_a$.
By assumption the map is expanding, $|f_a'|>1$. Hence
the diagonal terms drop off exponentially, as $1/|f_a'|^{m+1}$,
the terms below the diagonal fall off even faster, and
we are justified in truncating $\Bmat{a}$, as truncating the
matrix to a finite one introduces only exponentially small errors.
In the local matrix approximation the traces of \evOper s are approximated by
\beq
\left. \tr {\Lop}^n\right|_{\mbox{\tiny saddles}}
=\sum_{p} n_p \sum_{r=1}^{\infty} \delta_{n, n_p r} \tr \Lmat{p}^r
= \sum_{j=0}^{\infty}C_{n j}\sigma^j
\,,
\label{tracenp}
\eeq
where
$\tr \Lmat{p} = \tr{ \Lmat{\cl{p}}\Lmat{2}\cdots \Lmat{1}}$
is the contribution of the $p$ cycle, and the power series in
$\sigma^j$ follows from the expansion \refeq{BtoL} of $\Lmat{a}$
in terms of $\Bmat{a}$. The subscript
{\tiny saddles} is a reminder that this is a saddle-point
approximation to $\tr {\Lop}^n$ (see \refref{noisy_Fred} for a discussion),
valid as an asymptotic series in the limit of weak noise.
As a simple check of the above formulas, consider the noiseless case,
for which the $(\Lmat{a})_{m'm} = (\Bmat{a})_{m'm}$ matrices are
a representation of the deterministic \FPoper\ $\left. \Lop \right|_{\sigma=0}$.
The $\Lmat{a}$ are triangular with diagonal elements
$
(\Lmat{a})_{m m}= \frac{1}{|f_a'|f_a'^{m}}
\,.
$
The trace of the $\Lop$ on a periodic orbit $p$ is therefore
\[
\tr \Lmat{p}
= \tr{ \Lmat{\cl{p}}\Lmat{2}\cdots \Lmat{1}}
=\sum_{m=0}^{\infty} \frac{1}{|\ExpaEig_p|\ExpaEig_p^{m}}
=\frac{1}{|1-\ExpaEig_p|}
\,,
\]
and we recover the standard
deterministic trace formula\rf{QCcourse} for the {\FPoper}
\beq
\tr{\Lop}^n=
\sum_p \cl{p} \sum_{r=1}^\infty
\delta_{n,\cl{p} r}
\frac{1}{|1-\ExpaEig_p^r|}
\,.
\eeq
\section{Perturbative corrections to eigenvalues}
\label{trtoeig}
The eigenvalue condition \refeq{eigCond}
is an implicit equation for the eigenvalue $\eigenvL=\eigenvL(\sigma)$
of form $\eigCond(\sigma,\eigenvL(\sigma)) = 0$.
As the eigenvalue condition is satisfied for any $\sigma$,
all total derivatives of the eigenvalue condition with respect to $\sigma$
vanish, leading to
\bea
0 &=&
{d \over d\sigma} \eigCond(\sigma,\eigenvL(\sigma ))
=
{d \eigenvL \over d\sigma} {\pde \eigCond \over \pde \eigenvL}
\,+\,
{\pde \eigCond \over \pde \sigma}
\continue
0 &=&
{d^2 \eigenvL \over d\sigma^2}
{\pde \eigCond \over \pde \eigenvL}
\,+\,
\left({d \eigenvL \over d\sigma}\right)^2
{\pde^2 \eigCond \over \pde \eigenvL^2}
\,+\, 2 {d \eigenvL \over d\sigma}
{\pde^2 \eigCond \over \pde \sigma \pde \eigenvL}
\,+\,{\pde^2 \eigCond \over \pde \sigma^2}
\label{sec-derPert} \\
0 &=&
{d^3 \eigenvL \over d\sigma^3}
{\pde \eigCond \over \pde \eigenvL}
\,+\,
3{d^2 \eigenvL \over d\sigma^2}
{d \eigenvL \over d\sigma}
{\pde^2 \eigCond \over \pde \eigenvL^2}
\,+\,
\left({d \eigenvL \over d\sigma}\right)^3
{\pde^3 \eigCond \over \pde \eigenvL^3}
\ceq
\,+\, 3 {d^2 \eigenvL \over d\sigma^2}
{\pde^2 \eigCond \over \pde \sigma \pde \eigenvL}
\,+\, 3 \left({d \eigenvL \over d\sigma}\right)^2
{\pde^3 \eigCond \over \pde \sigma \pde \eigenvL^2}
\continue
\ceq \,+\, 3 {d \eigenvL \over d\sigma}
{\pde^3 \eigCond \over \pde \sigma^2 \pde \eigenvL}
\,+\,{\pde^3 \eigCond \over \pde \sigma^3}
\,,
\nnu
\eea
and so on.
$\eigenvL(0)$ can be computed by cycle expansions for a deterministic,
noiseless flow. $\sigma \neq 0$ then parametrizes
a weak perturbation to the deterministic
\FPoper\ $\left. \Lop \right|_{\sigma=0}$.
The above formulas enable us to compute recursively, order by order in
$\sigma^n$, the perturbative corrections to the eigenvalues of $\Lop$
\beq
\eigenvL(\sigma)
= \sum_{m=0}^\infty \eigenvL_m \sigma^m
\,,\qquad
\eigenvL_m = {1 \over m!} \left.{d^m~ \over d\sigma^m}\eigenvL(\sigma) \right|_{\sigma=0}
\,,
\ee{pertExp1}
in terms of partial derivatives of the eigenvalue condition
$\eigCond(\sigma,\eigenvL(\sigma))$
\beq
F_{kl} = \left.{\pde^{k+l}~~~ \over \pde \eigenvL^k \pde \sigma^l}
\eigCond(\sigma,\eigenvL)
\right|_{\sigma=0, \eigenvL = \eigenvL(0)}
\,.
\eeq
In this notation the formulas \refeq{sec-derPert} for $\eigenvL_m$
take the form
\bea
\eigenvL_1 &=& - {F_{01} \over F_{10}}
\continue
\eigenvL_2 &=& -{1 \over 2 \, F_{10}}\left( F_{02} + 2 \, F_{11} \, \eigenvL_1 + 2 \, F_{20} \, \eigenvL_1^{2} \right)
\label{EigPert-Diag}\\
\eigenvL_3 &=& -{1 \over 3! \, F_{10}}
\left(
F_{01} + 3 \, F_{12} \, \eigenvL_1 + 6 \, F_{11} \, \eigenvL_2
\right.
\ceq \qquad\qquad
\left.
+ \, 3 \, F_{21} \, \eigenvL_1^{2} + 6 \, F_{20} \, \eigenvL_1 \, \eigenvL_2 + F_{30} \, \eigenvL_1^{3} \right)
\,.
\nnu
\eea
As shown in \refref{QCcourse}, $F_{kl}$
can be computed from explicit cycle expansions.
However, in numerical calculations
we find it more expedient to proceede by first expressing
the {\fd} $F$ in terms of the cumulants.
The traces of $\Lmat{}^n$ evaluated by \refeq{BtoL} yield a series in
$\sigma^j$, and
the $\sigma^j$ coefficients $Q_{n j}$ in the cumulant expansion
\beq
F=\det(1-z\Lop) = 1-\sum_{n=1}^{\infty}\sum_{j=0}^{\infty}Q_{n j}z^n\sigma^j
\label{qum}
\eeq
are then obtained recursively from the traces, as in \refeq{Fd-cyc-exp-Diag}:
\beq
Q_{n j} = \frac{1}{n}\left(C_{n j}
- \sum_{k=1}^{n-1}\sum_{l=0}^{j}Q_{k,j-l}C_{n-k,l}
\right)
\,.
\ee{Q-Sig-expan}
This gives $F = F(z = 1/\eigenvL \, , \sigma)$ and the partial derivatives $F_{kl}$ can be found.
Substituted in \refeq{EigPert-Diag} they yield the perturbative corrections to the eigenvalues. The above calculations can be efficiently done by manipulating formal Taylor series.
\section{Numerical tests}
Here we continue the calculations of the eigenvalue corrections
described in \refrefs{noisy_Fred,conjug_Fred},
where more details and discussion may be found.
We test our perturbative expansion on the repeller
of the 1-dimensional map
\beq
f(x)=20\left(\frac{1}{16}-\left(\frac{1}{2}-x\right)^4\right)
\,.
\ee{testQuartic}
This repeller is a clean example of an ``Axiom~$A$'' expanding
system of bounded nonlinearity and complete binary symbolic dynamics,
for which the deterministic \evOper\ eigenvalues
converge super-exponentially with the cycle length\rf{Rugh92}.
We start the numerical calculations by determining
all prime cycles up to a given length.
For each prime cycle $p$ we compute
the truncated evolution matrix $\Lmat{p}$ and its repetitions $\Lmat{p}^r$
to the given order in $\sigma$, and evaluate the traces \refeq{tracenp}.
For the map at hand we find that truncations
of size [$16\times 16$] suffice to achive double precision accuracy
for most cycles. However, as the short orbits are less unstable, they require
larger matrix truncations in order to attain the same precision, and
we employ a [$28\times 28$] truncation for the $2$-cycles, and a
[$34\times 34$] truncation for the fixed points.
With the coefficients in the traces expansion \refeq{tracenp}
evaluated numerically, the cumulants and the perturbative eigenvalue
corrections follow from \refeq{Q-Sig-expan} and \refeq{EigPert-Diag}.
In case at hand, a good first approximation is obtained already at $n=2$
level, using only 3 prime cycles,
and $n=6$ (23 prime cycles in all) is in this example sufficient
to exhaust the limits of double precision arithmetic.
\begin{figure}[hbt]
\centerline{\psfig{figure=qum1.eps,width=9cm}}
\caption{The perturbative corrections
\refeq{Q-Sig-expan} to the cumulants $Q_{n j}$ plotted as a function
of cycle length $n$ (for perturbation orders $j=0,2,4,6,8$) all exhibit
super-exponential convergence.}
\label{qumfig}
\end{figure}
The size of the cumulants is indicated in \reffig{qumfig},
and the perturbative corrections to the leading eigenvalue
of the weak-noise \evOper\ are given in \reftab{tabPert}.
Encouragingly, the value of
$\eigenvL_6 = 2076.47\ldots$
computed here is not wildly different to our previous numerical
estimate\rf{conjug_Fred} of
2700.
Both the cumulants and the eigenvalue corrections exhibit a
super-exponential convergence with the truncation cycle length $n$.
The super-exponential convergence has been proven for the deterministic,
$\eigenvL_0$ part of the eigenvalue\rf{Rugh92}, but the proof
has not been extended to the stochastic evolution operators.
We have chosen to test the formalism on
this simple example, as here we are in a fortunate situation that
the escape rate for arbitrary noise strength $\sigma$ can be
calculated numerically by other methods to a rather high accuracy.
For example, one can discretize the stochastic kernel on a
spatial lattice\rf{noisy_Fred} and determine numerically
the leading eigenvalue.
The perturbative result in terms of periodic orbits and the weak noise corrections
is compared to the eigenvalue computed by the numerical
lattice discretizationin \reffig{nufig2}, with the absolute difference between
the numerical and the $m$th order
perturbative results plotted.
We see that the perturbative result
$\eigenvL(m,\sigma)=\sum_{k=0}^{m/2} \eigenvL_{2k}\sigma^{2k}$
indeed improves as more perturbative terms are added.
\begin{figure}[hbt]
\centerline{\psfig{figure=nufig2.eps,width=9cm}}
\caption{
The difference between the numerical and perturbative eigenvalue
$|\eigenvL(\sigma) - \eigenvL(m,\sigma)|$.
The plateau at $10^{-7}$ is a numerical artifact
due to the limited accuracy of the lattice discretization calculation.
\label{nufig2}}
\end{figure}
\section{Summary and outlook}
In this paper we study evolution of a classical dynamical system
with additive noise. In the limit of weak noise the traces of
the corresponding {\evOper} are approximated by sums of local
traces computed on periodic orbits.
Here we present a new, computationally efficient technique for
evaluation of these local traces based on
a matrix representation of the {\evOper}, and show that method is
powerful enough to enable us to
compute 2 more orders of perturbation theory.
The local matrix representation can be interpreted as follows.
Substituting \refeq{tracenp} into \refeq{det-tr-Diag} we obtain
\bea
\left.
\det(1 - z \Lop)
\right|_{\mbox{\tiny saddles}}
=
\prod_{p} \det(1 - z^{n_p} \Lmat{p})
\,.
\label{fredholmprod}
\eea
In other words, in
the saddle-point approximation the spectrum of the {\em global}
{\evOper} $\Lop$
is in this approach
pieced together from the {\em local} spectra computed cycle-by-cycle
on neighborhoods of individual prime cycles with periodic boundary
conditions. Vattay\rf{VatBS} was first to formulate the $\hbar$
corrections to the semi-classical Gutzwiller theory in terms of local
spectra.
Here we have shown that also the stochastic flows can be suspended
on the skeleton of classical periodic orbits in this way.
With so many orders of perturbation theory, we are now
poised to address the issues raised by the asymptotic series nature
of perturbative expansions. We can now hope to resum the series to all
orders, making use of techniques such as the Borel resummation, the asymptotic
expansions of general integrals of saddlepoint type, and
asymptotics beyond all orders\rf{Dingle}.
All of this is beyond the scope of
the present paper, and we defer a full discussion of asymptotics
to a forthcoming paper\rf{asym_Fred}.
\section{Acknowledgements}
G.V. and G.P. gratefully acknowledges the financial support of
the Hungarian Ministry of Education, FKFP 0159/1997, OMFB, OTKA T25866/F17166.
G.V. thanks
Bruno Eckhardt the cordial hospitalty at the Department of Physics of
the Philipps-Universit\"at Marburg and the Humboldt Fundation for
support.
G.P. thanks the EU network ``Pattern formation, noise and
spatio-temporal chaos in complex systems'', TMR contract
ERBFMRXCT960085, for partial
support. N.S. is supported by the Danish Research Academy Ph.D.
fellowship.
|
\section{Overview of the Cosmic Microwave Background}
In 1948, Alpher \& Hermann \cite{AlpHer49} realized that if
light elements were produced in a hot big bang, as Gamow and
others had suggested \cite{Gam46}, then the Universe today
should have a temperature of about 5 K. When Penzias
\& Wilson discovered an anomalous background in 1964,
consistent with a blackbody spectrum at a temperature of $\sim3$
K \cite{PenWil65}, Dicke and his
collaborators immediately recognized it as the radiation
associated with this nonzero cosmological temperature
\cite{Dicetal65}. Subsequent observations that confirm a
remarkable degree of isotropy (apart from a dipole
\cite{CorWil76,SmoGorMul77}, which can be interpreted as
our motion of $627\pm22$ km~s$^{-1}$ with respect to the
blackbody rest frame
\cite{Smoetal91,Smoetal92,Kogetal93,Fixetal94}) suggest an
extragalactic origin for this cosmic microwave background
(CMB). Strong upper limits to any angular cross-correlation
between the CMB temperature and the extragalactic X-ray
background intensity \cite{Kneetal97,BouCriTur98} suggest that the CMB
comes from redshifts greater than those ($z\simeq2-4$)
probed by
the active galactic nuclei and galaxy clusters that produce the
X-ray background. This evidence, as well as the
exquisite blackbody spectrum
of the CMB \cite{Matetal94,Wrietal94,Fixetal96}, further supports
the notion that this radiation is the cosmological blackbody
postulated by Alpher \& Hermann.
Although they have a Planck spectrum, CMB photons are not in
thermal equilibrium. The mean free path for scattering of
photons in the Universe must be huge, or else we would not see
galaxies and quasars out to distances of thousands of
Mpc.\footnote{Mpc$=3.3\times10^6$~light years$=3.09\times10^{24}$~cm.}
So where did these photons come from? At early times ($t\la10^5$ y;
redshifts $z\ga1000$), the temperature of the Universe exceeded
an eV,
so the Universe consisted of a plasma of free electrons
and light nuclei. CMB photons were tightly coupled to this plasma via
Thomson scattering from the free electrons. At a redshift of
$z\simeq1000$, the temperature dropped below a few eV, and
electrons and nuclei combined to form atoms. At this point,
photons ceased interacting.
A detailed analysis of ``recombination'' and
the almost simultaneous (although slightly later) decoupling of
photons shows that CMB photons last scattered near a redshift of
$z\simeq1100$ \cite{JonWys76,KolTur90,Pee93}.
When we look at these CMB photons coming to us from all
directions in
the sky, we are therefore looking directly at a
spherical surface in the Universe that surrounds us at a
distance of $\sim10^4$ Mpc, as it was when the
Universe was only about 300,000 years old. The temperature of
the CMB is found to be the same, to roughly one part in $10^5$, in
every direction on the sky. This remarkable
isotropy poses a fundamental
conundrum for the standard big-bang theory. When these photons
last scattered, the size of a causally connected region of the
Universe was roughly 300,000 light years, and such a region
subtends an angle of only one degree on the sky. Thus, when we
look at the CMB, we are looking at roughly 40,000 causally disconnected
regions of the Universe. How is it, then, that each of these
has the same temperature to one part in $10^5$? This is the
well-known isotropy, homogeneity, or horizon problem.
Another fundamental question in cosmology today is the origin
of the large-scale structure of the galaxy distribution.
The simplest and most plausible
explanation is that the observed inhomogeneities grew from tiny density
perturbations in the early Universe via gravitational instability.
Mass from underdense
regions is drawn towards overdense regions, and in this way,
small primordial perturbations are amplified into the
structure we see in the
Universe today. New support for this hypothesis was provided by
the Cosmic Background Explorer (\COBE) detection of temperature
differences in the CMB of
roughly one part in $10^5$ \cite{Smoetal90}.
Heuristically, density perturbations induce
gravitational-potential perturbations at the surface of last
scatter; photons that arrive from denser regions
climb out of deeper potential wells and thus appear redder
than those from underdense regions (the Sachs-Wolfe effect
\cite{SacWol67}). Thus, the temperature fluctuations seen with
\COBE\ provide a snapshot of the tiny primordial perturbations
that gave rise to the large-scale structure we see in the
Universe today. But this raises a second question: If
large-scale structure grew via gravitational infall from tiny
inhomogeneities in the early Universe, where did these
primordial perturbations come from?
Before \COBE, there was no shortage of ideas for the origin of
large-scale structure, and, quite remarkably, all causal
mechanisms for producing
primordial perturbations have come from new ideas in particle theory:
primordial adiabatic perturbations from inflation
\cite{GutPi82,Haw82,Lin82b,Sta82,BarSteTur83},
late-time phase transitions \cite{Was86,HilSchFry89}, a
loitering Universe \cite{SahFelSte92}, scalar-field ordering
\cite{Vil82,Pre80}, topological defects \cite{ZelKozOku74,Kib76}
(such as cosmic strings \cite{Zel80,Vil81,SilVil84,Tur85},
domain walls \cite{HilSchFry89,PreRydSpe89}, textures
\cite{Tur89,Tur91}, or global monopoles \cite{BarVil89,BenRhi91}),
superconducting cosmic strings \cite{Wit85,WitThoOst86},
isocurvature axion perturbations
\cite{SecTur85,TurWil91,AxeBraTur83,SteTur83,KofLin87,Lyt90,Lin91},
etc.
However, after \COBE, primordial
adiabatic perturbations (perturbations to the total density with
equal fractional number-density perturbations in each species in the
Universe) seem to provide the only workable models. Such
perturbations are produced naturally during inflation, a period
of exponential expansion in the early Universe driven by the
vacuum energy associated with some new scalar field
\cite{Gut81,Lin82a,AlbSte82}.
With adiabatic
perturbations, hotter regions at the surface of last scatter are
embedded in deeper potential wells, so the reddening due to
the gravitational redshift of the photons from these regions
partially cancels the higher intrinsic temperatures.
When normalized to the amplitude of density perturbations indicated by galaxy
surveys, alternative models generically produce a larger
temperature fluctuation than that measured by \COBE\
\cite{KodSas86,EfsBon87,JafSteFri94}. Recently, more detailed
calculations of the expected CMB-anisotropy amplitude have led
proponents of topological defects, the primary
alternative to inflation, to concede
that these models have difficulty accounting for the origin of
large-scale structure \cite{PenSelTur97,AlbBatRob97,Alletal97}.
Although inflation now seems to provide
the best candidate for the
origin of large-scale structure,
the primary attraction of
inflation was originally that it
provided (and still does) the best (if not only) solution of
the horizon problem.
For these reasons, inflation
has taken center stage in cosmology. Although
inflation was for a long time speculative physics beyond the
realm of experimental tests, we are now entering a new era in
which the predictions of inflation will be tested with
unprecedented precision by CMB measurements.
The primary focus of this article is therefore to review
the predictions of inflation and how they will be tested with
the CMB. Although inflation currently seems to provide the most
promising paradigm for the origin of large-scale structure, it
is not yet well established. Moreover, although the simplest
topological-defect models seem to be ruled out, it is still
certainly plausible that some more involved models may be
able to account for large-scale structure. We therefore review
the CMB predictions of topological-defect models. We also
discuss a number of other promising links between the CMB
and particle physics that do not necessarily have to do with the
origin of structure, e.g.\ dark matter, neutrino properties,
decaying
particles, cosmological magnetic fields from early-Universe
phase transitions, parity violation, gravity theories, time
variation of fundamental parameters, and baryogenesis scenarios.
We are unfortunately unable to cover the larger bodies of
excellent work on the CMB in general, nor on the intersections
between particle physics and cosmology more generally.
Fortunately, a number of excellent reviews cover those subjects,
to which we cannot do justice here.
Lyth and Riotto \cite{LytRio99} review particle-physics models of
inflation; Liddle and Lyth discuss structure formation
in inflation-inspired cold-dark-matter models.
Lidsey et al \cite{Lidetal97} review the production of density
perturbations and reconstruction of the inflaton potential from
the power spectra of density perturbations and gravitational
waves. White, Scott, and Silk \cite{WhiScoSil94} review the CMB and structure
formation, and Hu and White \cite{HuWhi97b} provide a brief review of
the theory of CMB polarization. Finally, see
References \cite{SunZel80,Rep95,Bir99} for reviews of the Sunyaev-Zeldovich
effect (the scattering of CMB photons from hot gas in clusters
of galaxies), an intriguing and potentially very important
probe of the physics of clusters.
\section{Cosmic Microwave Background Observables}
\label{sec:cmbobservables}
\subsection{The Frequency Spectrum}
\label{sec:freqspectrum}
Standard cosmology predicts the CMB frequency spectrum to be
that of a perfect blackbody,
\begin{equation}
S(\nu;T) = {2 h c^2 \nu^3 \over e^x-1},
\label{eq:blackbody}
\end{equation}
where $x=hc\nu/kT$, $h$ is Planck's constant, $c$ is the
velocity of light, $\nu$ is the frequency, $k$ is Boltzmann's
constant, and $T$ is the temperature. Of the infinitude of
possible distortions to this spectrum, two common forms often
considered in the literature---Bose-Einstein and Compton
distortions---could arise from basic physical processes
before recombination.
If photons are released into the Universe from some nonthermal
process (e.g.\ decay of a massive particle) when the temperature
of the Universe exceeds roughly 1 keV
(redshifts $z\ga10^6$ when
the age of the Universe is $t\la10^7$ sec), they will come into
complete thermal equilibrium with the photons in the primordial
plasma. More precisely, they attain kinetic equilibrium through
Compton scattering, double Compton scattering, and bremsstrahlung,
and they attain
chemical equilibrium (chemical potential $\mu=0$) because the
rate for photon-number-changing processes (e.g.\
$\gamma\gamma\rightarrow\gamma\gamma\gamma$) that maintain a chemical
potential $\mu=0$ exceeds the expansion rate. Therefore,
if any electromagnetic
energy is released into the Universe at such early times, it
will have no observable effect on the CMB. However, if photons
are released at later times (but still before recombination),
they can distort the CMB frequency spectrum
\cite{SunZel80,DanDeZ77,Lig81,DanDeZ82,SarCoo83,FukKaw90,BurDanDeZ91a,BurDanDeZ91b,HuSil93b}.
\subsubsection{Bose-Einstein distortion}
Nonthermal photons produced in the redshift range $10^5
\mathrel{\mathpalette\fun <} z \mathrel{\mathpalette\fun <} 3\times10^6$ (temperatures $T\simeq 0.1-1$ keV and
ages $t\simeq10^{7-9}$ sec) can still attain kinetic
equilibrium, but they will not attain chemical equilibrium, as
interactions that change the photon number occur less rapidly
than the expansion rate. If electromagnetic energy
is released at these times, the CMB frequency dependence will be
that of a Bose-Einstein gas with a nonzero chemical potential,
\begin{equation}
S_\mu(\nu;T,\mu) = {2 h c^2 \nu^3 \over e^{x+\mu}-1},
\label{eq:boseeinstein}
\end{equation}
where $\mu$ is the (dimensionless) chemical potential. The
Far Infrared Absolute Spectrophotometer (FIRAS) result for $\mu$
is $\mu=-1\pm10\times10^{-5}$ or a 95\%
confidence-level upper limit of $|\mu| < 9\times10^{-5}$
\cite{Fixetal96}. It is possible that values of $\mu$ as small
as $10^{-6}$ could be probed by a future satellite mission
\cite{Shaetal95}.
\subsubsection{Compton distortion}
If photons are released at later times ($z\la10^5$) but still
before recombination ($z\simeq1100$; temperatures $T\simeq 1-100$
eV and times $t\simeq10^{9-13}$ sec), they do not have enough
time to come to either thermal or kinetic equilibrium and wind
up producing a ``Compton distortion'' of the form
\begin{equation}
S_y(\nu;T,y)={2 h c^2 \nu^3 \over e^x-1}\left(1 + y x {1
\over 1-e^{-x}} \left[x\coth(x/2)-4\right]\right),
\label{eq:comptony}
\end{equation}
to linear order in $y$
(the Kompaneets or ``Compton-$y$''
parameter). If some CMB photons were rescattered after
recombination by a hot intergalactic gas, this would also
produce a Compton-$y$ distortion. The FIRAS result for this
type of distortion is $y=-1\pm6\times10^{-6}$ or an upper limit
of $|y|<15\times10^{-6}$ at the 95\% confidence level
\cite{Fixetal96}. The
consensus among the experimentalists we have surveyed seems to
be that it would be difficult to improve on this limit.
\subsection{Temperature and Polarization Power Spectra}
The primary aim of forthcoming CMB satellite
experiments, such as NASA's Microwave Anisotropy Probe (MAP) \cite{MAP}
and the European Space Agency's Planck Surveyor \cite{Planck},
will be to map the temperature $T({\bf \hat n})$ of the CMB and its
linear polarization, described by Stokes parameters $Q({\bf \hat n})$
and $U({\bf \hat n})$, as functions of position ${\bf \hat n}$ on the sky.
Several temperature-polarization angular correlation
functions, or equivalently, power spectra, can be
extracted from such maps.
These quantities can be compared with detailed predictions from
cosmological models.
\subsubsection{Harmonic analysis for temperature anisotropies and
polarization}
{\it Temperature Anisotropies}\\
The temperature map can be expanded in spherical harmonics,
\begin{equation}
{T({\bf \hat n}) \over T_0} = 1 + \sum_{lm} a^{\rm T}_{(lm)}
Y_{(lm)}({\bf \hat n}),
\label{Texpansion}
\end{equation}
where the mode amplitudes are given by
\begin{equation}
a^{\rm T}_{(lm)}={1\over T_0}\int
d{\bf \hat n}\,T({\bf \hat n})\,Y_{(lm)}^*({\bf \hat n});
\label{Talms}
\end{equation}
this follows from orthonormality of the spherical harmonics.
Here, $T_0=2.728\pm0.002$ K is the cosmological mean CMB
temperature \cite{Fixetal96}.
{\it Linear polarization}\\
The Stokes parameters (where $Q$ and $U$ are measured with
respect to the polar ${\bf \hat\theta}$ and azimuthal ${\bf \hat
\phi}$ axes) are components of a $2\times2$ symmetric
traceless tensor with two independent components,
\begin{equation}
{\cal P}_{ab}({\bf \hat n})={1 \over 2} \left( \begin{array}{cc}
\vphantom{1\over 2}Q({\bf \hat n}) & -U({\bf \hat n}) \sin\theta \\
-U({\bf \hat n})\sin\theta & -Q({\bf \hat n})\sin^2\theta \\
\end{array} \right),
\label{eq:whatPis}
\end{equation}
where the subscripts $ab$ are tensor indices, and $Q$ and $U$
are given in temperature units.
Just as the temperature is expanded in terms of spherical
harmonics, the polarization tensor can be expanded
\cite{KamKosSte97a,KamKosSte97b,SelZal97,ZalSel97},
\begin{equation}
{{\cal P}_{ab}({\bf \hat n})\over T_0} =
\sum_{lm} \left[ a_{(lm)}^{{\rm G}}Y_{(lm)ab}^{{\rm
G}}({\bf \hat n}) +a_{(lm)}^{{\rm C}}Y_{(lm)ab}^{{\rm C}}({\bf \hat n})
\right],
\label{eq:Pexpansion}
\end{equation}
in terms of tensor spherical harmonics, $Y_{(lm)ab}^{\rm G}$
and $Y_{(lm)ab}^{\rm C}$. It is well known that a vector field
can be decomposed into a curl (C) and a curl-free (gradient) (G)
part.
Similarly, a
$2\times2$ symmetric traceless
tensor field can be decomposed into a tensor
analogue of a curl and a gradient part; the $Y_{(lm)ab}^{\rm G}$
and $Y_{(lm)ab}^{\rm C}$ form a complete orthonormal basis for the
``gradient'' (i.e.\ curl-free) and ``curl'' components of the
tensor field, respectively.\footnote{Our G and C are sometimes
referred to as the ``scalar'' and ``pseudo-scalar'' components
\cite{Ste96}, respectively, or with slightly different
normalization as E and B modes
\cite{ZalSel97} (although these should not be confused with the
radiation's electric- and magnetic-field vectors).} Lengthy but
digestible expressions for the
$Y_{(lm)ab}^{\rm G}$ and $Y_{(lm)ab}^{\rm C}$ are given in terms
of derivatives of spherical harmonics and also in terms of
Legendre functions in Reference \cite{KamKosSte97b}.
The mode amplitudes in Equation \ref{eq:Pexpansion} are given by
\begin{eqnarray}
a^{\rm G}_{(lm)}&=&{1\over T_0}\int d{\bf \hat n}\,{\cal P}_{ab}({\bf \hat n})\,
Y_{(lm)}^{{\rm G}
\,ab\, *}({\bf \hat n}),\cr
a^{\rm C}_{(lm)}&=&{1\over T_0}\int d{\bf \hat n}\,{\cal P}_{ab}({\bf \hat n})\,
Y_{(lm)}^{{\rm C} \,
ab\, *}({\bf \hat n}),
\label{eq:Amplitudes}
\end{eqnarray}
which can be derived from the orthonormality properties of these
tensor harmonics \cite{KamKosSte97b}.
Thus, given a polarization map ${\cal P}_{ab}({\bf \hat n})$, the G and
C components can be isolated by first carrying out the
transformations in Equation \ref{eq:Amplitudes} to the $a^{\rm
G}_{(lm)}$ and $a^{\rm C}_{(lm)}$ and then summing over the
first term on the right-hand side of Equation \ref{eq:Pexpansion}
to get the G component and over the second term to get the C
component.
\subsubsection{The power spectra}
\label{sec:powerspectra}
Theories for the origin of large-scale structure predict that
the mass distribution in the Universe is a single realization of
a statistically isotropic random field. In other words, the
Fourier components $\tilde\delta({\vec k})$ of the fractional
density perturbation $\delta({\vec x}) =
[\rho({\vec x})-\bar\rho]/\bar\rho$ [where $\rho({\vec x})$ is the
density at comoving position ${\vec x}$ and $\bar\rho$ is the
universal mean density] are random variables that have
expectation values $\VEV{\tilde\delta({\vec k})}=0$ and covariance
given by
\begin{equation}
\VEV{\tilde \delta({\vec k}) \tilde \delta({\vec k}')} = (2\pi)^3
\, \delta_D({\vec k}+{\vec k}') \, P_s(k).
\label{eq:scalarspectrum}
\end{equation}
Here $P_s(k)$ is the scalar power spectrum (so called because
density perturbations produce scalar perturbations to the
spacetime metric), or alternatively, the power spectrum for the
spatial mass distribution.
Statistical isotropy demands that
the power spectrum depends only on the amplitude (rather than
orientation) of ${\vec k}$.
Because the temperature perturbation and polarization of the CMB
are due to density perturbations, the $a_{(lm)}^{\rm X}$
must be random variables with zero mean, $\VEV{a_{(lm)}^{\rm
X}}=0$, and covariance,
\begin{equation}
\VEV{\left(a_{(l'm')}^{\rm X'} \right)^* a_{(lm)}^{\rm X}}
= C_l^{{\rm XX}'} \delta_{ll'}\delta_{mm'},
\label{eq:covariance}
\end{equation}
for ${\rm X},{\rm X}' = \{{\rm T,G,C}\}$. The statistical
independence of each $lm$ mode (i.e.\ the presence of the
Kronecker deltas) is a consequence of statistical
isotropy. The scalar spherical harmonics $Y_{(lm)}$ and the
gradient tensor spherical harmonics $Y_{(lm)}^{\rm G}$ have
parity $(-1)^l$, whereas the curl tensor spherical harmonics
$Y_{(lm)}^{\rm C}$ have the opposite parity, $(-1)^{l+1}$.
Thus, $C_l^{\rm TC}=C_l^{\rm GC}=0$ if the physics that gives rise
to temperature anisotropies and polarization is
parity-invariant. In this case, the two-point statistics of the CMB
temperature-polarization map are completely specified
by the four sets of moments, $C_l^{\rm TT}$, $C_l^{\rm TG}$,
$C_l^{\rm GG}$, and $C_l^{\rm CC}$. Nonzero $C_l^{\rm TC}$ or
$C_l^{\rm GC}$ would provide a signature of cosmological parity
breaking.
\subsubsection{Angular correlation functions}
The temperature two-point correlation function is
\begin{equation}
C^{\rm TT}(\alpha) = \VEV{ {\Delta T({\bf \hat m})\over T} {\Delta T({\bf \hat n})
\over T}}_{{\bf \hat m} \cdot {\bf \hat n}=\cos\alpha},
\label{eq:correlationfns}
\end{equation}
where the average is over all pairs of points on the sky
separated by an angle $\alpha$. It can be written in terms of
the temperature power spectrum (i.e.\ the $C_l^{\rm TT}$) as
\begin{equation}
C^{\rm TT}(\alpha) = \sum_l {2 l +1 \over 4 \pi}
C^{\rm TT}_l P_l(\cos\alpha),
\label{eq:corrpower}
\end{equation}
where $P_l(\cos\alpha)$ are Legendre polynomials. Likewise,
orthonormality of Legendre polynomials guarantees that the
multipole coefficients, $C_l^{\rm TT}$, can be written as
integrals over the product of the correlation function and a
Legendre polynomial. Thus, specification of the correlation function is
equivalent to specification of the power spectrum, and vice
versa. CMB theorists and experimentalists have now adopted the
convention of showing model predictions and presenting
experimental results as power spectra ($C_l$) rather than as
correlation functions, and we subsequently stick to this
convention. Auto- and cross-correlation
functions for the Stokes parameters and temperature-polarization
cross-correlation functions can also be defined and written in
terms of the polarization and temperature-polarization power
spectra \cite{KamKosSte97b}, but we do not list them here.
In practice, the temperature intensity (or polarization)
can never be determined
at a given point on the sky; it can only be measured by a
receiver of some finite angular resolution (referred to as a
``finite beamwidth''). Thus, the correlation function in
Equation~\ref{eq:corrpower} cannot be measured; one can only
measure a smoothed version. Likewise, a finite beamwidth
$\theta_{\rm fwhm}$ (at full-width half maximum) limits
determination of the power spectrum to multipole moments
$l\mathrel{\mathpalette\fun <} 200\,(\theta_{\rm fwhm}/1^\ifmmode ^\circ\, \else $^\circ\,$ \fi)^{-1}$.
The Differential Microwave Radiometer (DMR) experiment
\cite{Smoetal92} aboard \COBE\ produced the first map of the
temperature of the
CMB. The receivers also provided some information on the
polarization, but the sensitivity was not sufficient to
detect the signal expected in most cosmological models.
Measurements of the CMB intensity
were made at three different frequencies (31.5, 53, and 90 GHz)
near the blackbody peak to disentangle the possible contribution
of foreground contaminants (e.g.\ dust or synchrotron emission)
{}from the Galaxy, as these would have a frequency spectrum that
differs from a blackbody.
The DMR beamwidth was $7^\ifmmode ^\circ\, \else $^\circ\,$ \fi$, so the temperature power
spectrum was recoverable only for $l\mathrel{\mathpalette\fun <} 15$. MAP, scheduled for
launch in the year 2000, will map the sky with an angular
resolution better than $0.3^\ifmmode ^\circ\, \else $^\circ\,$ \fi$ ($l\la700$). MAP should have
sufficient sensitivity to see the polarization, although probably not
enough to map the polarization power spectra precisely (the
polarization is expected to be roughly an order of magnitude
smaller than the temperature anisotropy). The
Planck Surveyor, scheduled for launch around 2007, will map the
temperature and polarization with even finer angular
resolution (out to $l\la2000-3000$). Its sensitivity should
be sufficient to map the polarization power spectra expected
{}from density perturbations (discussed below) with good precision.
\subsection{Gaussianity}
Angular three-point and higher $n$-point temperature correlation
functions can be constructed analogously to the two-point
correlation functions in Equation~\ref{eq:correlationfns}.
Fourier analogs of higher-order correlation functions can
be defined. In particular, the temperature bispectrum
$B(l_1,l_2,l_3)$ is the $l$-space version of the temperature
three-point correlation function. It is defined by
\begin{equation}
\VEV{a^{\rm T}_{(l_1 m_1)} a^{\rm T}_{(l_2 m_2)} a^{\rm
T}_{(l_3 m_3)}} = \left( \begin{array}{ccc} l_1 & l_2 & l_3
\\ m_1 & m_2 & m_3 \\ \end{array} \right) B(l_1,l_2,l_3),
\label{eq:bispectrum}
\end{equation}
where the array is the Wigner $3j$ symbol.
This particular $m$ and $l$ dependence follows from
the assumption of statistical isotropy. Closely related
statistics include the skewness and kurtosis (respectively, the
three- and four-point correlation functions at zero lag)
\cite{ScaVit91,LuoSch93,Luo94} and higher cumulants \cite{FerMagSil97}.
As discussed further below,
inflationary models predict the primordial distribution of
perturbations to be perfectly (or very nearly) Gaussian.
Gaussianity dictates that all the odd-$n$ $n$-point correlation
functions vanish and that for even $n$, the higher $n$-point correlation
functions can be given in terms of the two-point correlation
function.
Numerous other statistical tests of CMB Gaussianity have also
been proposed, including (but not limited to) topology of
temperature contours \cite{Col88,Gotetal90,Kogetal96}
and the related Minkowski
functionals \cite{WinKos97,SchGor98}, temperature peak
statistics \cite{BonEfs87,VitJus87,Kogetal96}, Fourier space patterns
\cite{FerMag97,LewAlbMag99},
and wavelet analysis \cite{PanValFan98,HobJonLas98}.
\begin{figure}
\centerline{\psfig{file=potentia.ps,width=32pc}}
\caption{Two toy models for the inflationary potential.}
\vskip-12pt
\label{fig:potentials}
\end{figure}
\section{Predictions of Inflation}
\label{sec:inflationpredictions}
If the energy density of the Universe is dominated by matter or
radiation, then the expansion of the Universe is decelerating.
If so, the horizon grows more rapidly than the scale
factor. In such a Universe, objects that are now beyond our
horizon and therefore inaccessible to us will eventually enter
the horizon and become visible. Thus, the observable Universe
contains more information and is more complicated at later
times. Inflation postulates the existence of a period of
accelerated expansion in the early Universe. In such a
Universe, the scale factor grows more rapidly than the horizon.
Thus, objects currently visible to any given observer will
eventually exit that observer's horizon (in much the same
way that objects that fall into a black hole disappear when they
pass through the black hole's event horizon). A period of
accelerated expansion therefore makes the Universe simpler and
smoother.
As we now discuss, this accelerated expansion also generically
drives the observable Universe to be flat and provides a
mechanism for producing primordial density perturbations and
gravitational waves.
\subsection{Scalar-Field Dynamics}
\label{sec:scalarfields}
Inflation supposes the existence of some new scalar field $\phi$ (the
``inflaton''), with a
potential $V(\phi)$ that
roughly resembles either of those
shown in Figure \ref{fig:potentials}. The shape is not particularly
important. All we require is that, at some time in the
history of the Universe, the field is displaced from the minimum
of the potential, and then it rolls slowly
How slowly is slowly enough? This is determined by the Friedmann
equation,
\begin{equation}
H^2 \equiv \left( {\dot a \over a}\right)^2 ={8 \pi G\rho
\over 3} - {k\over a^2} = {8 \pi G
\over 3} \left({1\over2} \dot\phi^2 + V(\phi) \right) -{k
\over a^2},
\label{eq:friedmann}
\end{equation}
which governs the time $t$ dependence of the scale factor
$a(t)$ of the Universe (the dot denotes derivative with
respect to time), as well as the scalar-field equation of motion,
\begin{equation}
\ddot \phi + 3H \dot\phi +V'(\phi)=0,
\label{eq:eom}
\end{equation}
where $V'=dV/d\phi$.
In Equation~\ref{eq:friedmann}, $\rho$ is the energy density of the
Universe, which is assumed to be dominated by the inflaton
potential-energy density $V(\phi)$ and kinetic-energy density
$\dot\phi^2/2$. The term
$k/a^2$ is the curvature term, and $k>0$, $k<0$, or $k=0$ for a
closed, open, or flat Universe, respectively. Note that the
expansion acts as a friction term for the scalar-field equation
of motion in Equation~\ref{eq:eom}. If
\begin{equation}
\epsilon\equiv {m_{\rm Pl}^2 \over 16 \pi} \left({ V' \over
V} \right)^2 \ll 1,
\label{eq:epsilon}
\end{equation}
and
\begin{equation}
\eta \equiv {m_{\rm Pl}^2 \over 8\pi} \left[{ V'' \over V}
- {1 \over 2} \left({V' \over V} \right)^2 \right] \ll 1,
\label{eq:eta}
\end{equation}
where $m_{\rm Pl}=1.22 \times 10^{19}$ GeV is the Planck mass,
then the field rolls slowly enough so that the requirement for
acceleration [$p<-\rho/3$, where $p= (1/2)\dot\phi^2-V$ is the
pressure and $\rho= (1/2)\dot\phi^2 +V$ is the energy
density] is satisfied. (Note that the definitions of $\epsilon$ and
especially of $\eta$ may differ in some papers.)
The identity of the inflaton remains a mystery. It was originally
hypothesized to be associated with a Higgs field in
grand unified theories,
but it may also have something to do
with Peccei-Quinn symmetry breaking, a dilaton field,
electroweak-symmetry breaking \cite{KnoTur93}, some new
pseudo-Nambu-Goldstone symmetry \cite{FreFriOli90,Adaetal93},
supersymmetry \cite{RanSolGut96}, or some other new physics. As
discussed below, the primary predictions of slow-roll inflation
do not depend on the details of the physics responsible for
inflation but rather on some gross features that are easily
quantified.
\subsection{The Geometry}
Given any inflationary potential $V(\phi)$, the equations of
motion in Equations \ref{eq:eom} and \ref{eq:epsilon} can be
solved numerically, if not analytically.
Heuristically, during inflation, the potential $V(\phi)$ is
roughly constant, and $\dot\phi^2 \ll V(\phi)$. If the curvature term is
appreciable initially, it rapidly decays relative to the
potential term as the Universe expands, and the solution for the
scale factor approaches an exponential, $a(t) \propto
e^{-Ht}$. If $k$ is nonzero initially, the curvature term is
then driven exponentially to zero during the inflationary epoch.
In other words, if the duration of inflation is sufficiently
long to place the observable Universe in a causally connected
pre-inflationary patch, then the curvature radius is generically
inflated to an exponentially (and unobservably) large value. In
the language above, any initial nonzero curvature disappears
beyond the horizon during accelerated expansion.
Thus, the first prediction of slow-roll inflation is that
the Universe should be flat today; i.e.\ the total density of
all components of matter should sum to the critical density.
\subsubsection{``Open inflation''}
It is, of course, mathematically possible that
inflation did occur but that the inflationary epoch was
prematurely
terminated \cite{LytSte90,KamSpe94} at just the right time so that the
Universe today would be open with density
$\Omega_0\simeq0.3$. Such a model requires some additional
mechanism (e.g.\ another prior period of inflation and/or some
arbitrary new ``feature'' in the inflaton potential) to solve the
isotropy problem as well as to produce density perturbations.
Several such open-inflation models have recently been constructed
\cite{Got82,LytSte90,BucGolTur95,RatPee95,Lin95,CorSpeSta96,HawTur98},
motivated by observations that suggest $\Omega_0\simeq0.3$.
The predictions of a scale-invariant spectrum and Gaussian
perturbations (discussed below) are the same as in
ordinary inflation, but the Universe would be open.
We do not find these models even nearly as compelling as
the ordinary slow-roll
models that produce a flat Universe, although some theorists may
disagree. Fortunately, the correct model will not be determined by
debate; forthcoming CMB measurements, described below,
should distinguish conclusively between these two classes---simple and
elegant versus complicated and unappealing---of inflationary models.
\subsection{Density Perturbations, Gravitational Waves, and the
Inflationary Observables}
\subsubsection{Production of density perturbations}
Density perturbations are produced as a result of novel
quantum-mechanical effects (analogous to the production of
Hawking radiation from black holes) that occur in a Universe
with accelerating expansion
\cite{GutPi82,Haw82,Lin82b,Sta82,BarSteTur83}.
This process has been reviewed in detail recently \cite{Lin90,Lidetal97},
so here we review the physics only heuristically.
Consider perturbations $\delta\phi({\vec x},t)$ (as a
function of comoving position ${\vec x}$) to the
homogeneous slowly rolling field $\phi(t)$. These perturbations
satisfy a massless Klein-Gordon equation, and the equation
of motion for each Fourier mode $\widetilde{\delta\phi}({\vec k})$
is that of a simple harmonic oscillator in an expanding Universe.
At sufficiently early times, when the wavelength of any given
Fourier mode is less than the Hubble radius $H^{-1}$, it undergoes
quantum-mechanical zero-point oscillations.
However, if the expansion is accelerating, then the
physical wavelength of this comoving scale grows faster than
the Hubble radius and eventually becomes larger than $H^{-1}$. At this point,
crests and troughs of a given Fourier mode can no longer
communicate, and the zero-point
fluctuation becomes frozen in as a classical perturbation
$\delta\phi({\vec x})$ to the scalar field. Because the inflaton
potential is not perfectly flat, this induces perturbations to
the density $\delta\rho({\vec x}) =(\partial V/\partial
\phi)\delta\phi({\vec x})$.
\subsubsection{Production of gravitational waves}
Tensor perturbations to the spacetime metric (i.e.\
gravitational waves) satisfy a massless Klein-Gordon equation.
A stochastic background of gravitational waves are
therefore produced in the same way as classical perturbations to
the inflaton are produced \cite{AbbWis84}.
Moreover, the power spectra for the
inflaton-field perturbations and for the tensor metric
perturbations should be
identical. The power spectrum of density perturbations is a little
different from that for gravitational waves because a density
perturbation is related to a
scalar-field perturbation by $\delta\rho = (\partial
V/\partial\phi)\delta\phi$. The production of scalar and tensor
perturbations depends only on the expansion rate during
inflation. If the expansion rate were perfectly constant during
inflation, it would produce flat scalar and tensor power
spectra, $P_s \propto k$ (the ``Peebles-Harrison-Zeldovich''
\cite{PeeYu70,Har70,Zel72} spectrum) and $P_t(k) \propto {\rm
constant}$.
\subsubsection{Inflationary observables}
A constant expansion rate is an oversimplification
because the field must in fact be rolling slowly down the
potential during inflation. Given any specific functional form
for the potential, it is straightforward, using the tools of
quantum field theory in curved spacetimes (see e.g.\
\cite{BirDav82}), to predict precisely the functional forms
of $P_s(k)$ and $P_t(k)$. Measurement of these power spectra
could then be used to reconstruct the inflaton potential
\cite{Lidetal97}. Since the field must be
rolling fairly slowly during inflation, a good approximation (in
most models) can be obtained by expanding about a constant
expansion rate. In this slow-roll approximation, the primordial
scalar power spectrum is
\footnote{Note that this is the
spectrum for the primordial perturbations. After the Universe
becomes matter dominated
at a redshift $z\simeq10^4$, density
perturbations grow via
gravitational infall, and the growth factor depends on the
wave number. Therefore, the power spectrum for matter today is
different from the primordial spectrum (it becomes $k^{-4}$ times
the primordial spectrum at large $k$), but it is straightforward
to relate the primordial and current power spectra.}
\begin{equation}
P_s = A_s k^{n_s},
\label{eq:scalarPs}
\end{equation}
and the primordial power spectrum for tensor perturbations is
\begin{equation}
P_t = A_t k^{n_t}.
\label{eq:tensorPt}
\end{equation}
The amplitudes $A_t$ and $A_s$ and power-law indices $n_s$ and
$n_t$ have come to be known as the ``inflationary
observables.'' These parameters can
provide information on the inflaton potential. In the slow-roll
approximation, the power-spectrum indices are roughly constant and given by
\cite{Sta85,LidLyt92,Davetal92,LucMatMol92,LidCol92,Tur93a,Adaetal93,Tur93b,Copetal93a,Copetal93b,Copetal94,LidTur94,Lidetal97}
\begin{equation}
n_s=1-4\epsilon+2\eta, \qquad n_t= -2\epsilon,
\label{eq:nsnt}
\end{equation}
where $\epsilon$ and $\eta$ are the slow-roll parameters given
in Equations \ref{eq:epsilon} and \ref{eq:eta}. Strictly
speaking, $\epsilon$ and $\eta$ may change (logarithmically with
$k$) during inflation \cite{KosTur95,Lidetal97}, but, as the name
implies, the field rolls slowly during slow-roll
inflation, so the running of the spectral indices is, for all
practical purposes, very small.
The amplitudes $A_s$ and $A_t$ are similarly fixed by the
inflaton potential, but their precise values depend on
Fourier conventions and on how the scale factor today is
chosen. However, $A_s$ and $A_t$ are proportional, respectively, to the
amplitude of the scalar and tensor contributions
to $C_2^{\rm TT}$, the quadrupole moment of the CMB temperature.
In terms of the slow-roll parameter $\epsilon$ and height $V$ of the
inflaton potential during inflation, these CMB observables are
\begin{eqnarray}
{\cal S} &\equiv & 6\, C_2^{{\rm TT},{\rm scalar}}=
0.66\, {V \over \epsilon m_{\rm Pl}^4}
\nonumber\\
{\cal T} &\equiv & 6\, C_2^{{\rm TT},{\rm tensor}}= 9.2 {V
\over m_{\rm Pl}^4}.
\label{eq:amplitudes}
\end{eqnarray}
For nearly scale-invariant spectra, \COBE\ fixes $C_2^{\rm
TT}=C_2^{\rm TT,scalar}+C_2^{\rm
TT,tensor}=(1.0\pm0.1)\times10^{-10}$.
In terms of the slow-roll parameters, the tensor-to-scalar ratio
is usually defined to be
\begin{equation}
r \equiv { {\cal T} \over {\cal S}} = 13.7\,\epsilon.
\label{eq:TtoS}
\end{equation}
Comparing Equation~\ref{eq:TtoS} with Equation~\ref{eq:nsnt}, we
observer that the observables
$n_t$ and $r$ must satisfy a consistency relation, $n_t=-0.145
r$, in slow-roll models.
To summarize, slow-roll inflation models (which account for the
overwhelming majority of inflation models that appear in the
literature) are parameterized by (\textit{a}) the height $V$ of the
inflaton potential (i.e.\ the energy scale of inflation), (\textit{b})
$\epsilon$, which depends on the first derivative $V'$ of the
inflaton potential, and (\textit{c}) $\eta$, which depends additionally on
the second derivative $V''$.
The discussion above suggests that because the inflaton is always
rolling down the potential, the scalar spectral index must be
$n_s<1$. Although this may be true for simple single-field
inflation models, more complicated models (e.g.\ with multiple
fields or with different potentials) may produce ``blue''
spectra with $n_s>1$ \cite{MolMatLuc94}.
\COBE\ alone already constrains $V^{1/4} \leq 2.3\times10^{16}$
GeV. With some additional (but reasonable) modeling, the \COBE\
constraint can be combined with current degree-scale
CMB-anisotropy measurements and large-scale-structure
observations to reduce this to $V^{1/4} \leq 1.7\times10^{16}$
GeV (e.g.\ \cite{ZibScoWhi99}).
The \COBE\ anisotropy implies $n_s=1.1\pm0.3$
if it is attributed entirely to density perturbations
\cite{Benetal96}, or $n_t=0.2\pm0.3$ if it is
attributed entirely to gravitational waves. Therefore, barring
strange coincidences, the \COBE\ spectral index and relations
above seem to suggest that if slow-roll inflation is right, then
the scalar and tensor spectra must both be nearly scale-invariant
($n_s\simeq 1$ and $n_t\simeq 0$).
\subsection{Character of Primordial Perturbations}
\subsubsection{Adiabatic versus isocurvature}
The density perturbations
produced by quantum fluctuations in the inflaton field are
referred to as adiabatic, curvature, or
isentropic perturbations.
These are perturbations to the
total density of the Universe, or equivalently, scalar
perturbations to the spacetime metric. Adiabaticity further
implies that the spatial distribution of each species in the
Universe (e.g.\ baryons, photons, neutrinos, dark matter) is the
same---that is, the ratio of number densities of any two of
these species is everywhere the same.
Adiabatic perturbations can be contrasted with primordial
isocurvature, or equivalently, pressure or entropy
perturbations, perturbations to the ratios between the various
species in the Universe (usually in a Universe with a
homogeneous total density). Such varying ratios would set up
perturbations to the pressure or equivalently to the entropy.
When two initially causally-disconnected regions with different
pressures come into causal contact, the pressure perturbations
push matter around, thus seeding large-scale structure.
{\it Axion Inflation}\\
Although adiabatic perturbations are generically produced
during inflation, it is also possible to obtain isocurvature
perturbations. One example is
isocurvature perturbations to an axion density from quantum
fluctuations in the Peccei-Quinn field during inflation
\cite{SecTur85,TurWil91,AxeBraTur83,SteTur83,KofLin87,Lyt90,Lin91}.
As discussed below, comparison of the measured amplitude of
CMB anisotropies with the amplitude of galaxy clustering
essentially rules out these models. Inflation models
that produce both adiabatic and isocurvature perturbations
have also been considered
\cite{MukZel91,PolSta92,PetPolSta94,StaYok95,GarWan96,SasSte96,KawSugYan96,LinMuk97,MukSte98,Kanetal98a,Kanetal98b};
future experiments should tightly constrain the relative
contributions of the two types of perturbations.
\subsubsection{Causal versus acausal}
Perturbations produced by inflation are said to be
``super-horizon'' or ``acausal.'' This simply refers to the
fact that inflation produces a primordial (meaning
before matter-radiation equality, when gravitational
amplification of perturbations can begin) spectrum of
perturbations of all wavelengths, including those much
larger than the Hubble length at any given time. This is to be
contrasted, for example, with ``causal'' models of structure
formation, in which perturbations are generated by
causal physics on scales smaller than the horizon.
Since inflation implies distance scales much
larger than the Hubble length can be within a causally connected
pre-inflationary patch, the term acausal is really a
misnomer.
\subsubsection{(Nearly) Gaussian distribution of perturbations}
If the inflaton potential is flat enough for the slow-roll
approximation to be valid, then each Fourier mode of the
inflaton perturbation evolves independently; that is, the
inflaton behaves essentially like an uncoupled massless
scalar field. As a result, inflation predicts that the
primordial density field is a realization of a Gaussian random
field: each Fourier mode is decoupled from every other, and the
probability distribution for each is Gaussian.
Of course, Gaussianity is an approximation that becomes
increasingly valid in the slow-roll limit, in which
the inflaton perturbation can be treated as a
noninteracting scalar field.
Deviations from
Gaussianity are generally accepted to be small, and most
theorists have adopted a pure Gaussian distribution as a
prediction of inflation. However, the deviations in some models
might be observable, and if so, would shed light on the physics
responsible for inflation
\cite{AllGriWis87,SalBonBar89,Sal92,FalRanSre93,Ganetal94,Gan94}.
This can be quantified more precisely with the three-point
statistic \cite{FerMagGor98},
\begin{equation}
I_l^3 \equiv {1 \over (2l+1)^{3/2} (C_l^{\rm TT})^{3/2}}
\left( \begin{array}{ccc} l & l & l
\\ 0 & 0 & 0 \\ \end{array} \right) B(l,l,l).
\end{equation}
In slow-roll models with smooth inflaton potentials, the
prediction for this quantity is (L Wang, M Kamionkowski,
manuscript in preparation)
\begin{equation}
\sqrt{l(l+1)}I_l^3 = {2 \over m_{\rm Pl}^2} \sqrt{ 3V
\over \epsilon} (3\epsilon -2\eta).
\label{eq:Il3prediction}
\end{equation}
Thus, in slow-roll models, one expects $I_0 \mathrel{\mathpalette\fun <} 10^{-6}$ (unless
for some unforeseen reason $\epsilon$ is extremely small and
$\eta$ is not), too small to be observed. A larger non-Gaussian
signal may conceivably arise if there is a glitch in the
inflaton potential, but even this non-Gaussianity would be
extremely small (L Wang, M Kamionkowski, manuscript in
preparation). Detection of nonzero $I_0$ would thus rule out
the simplest slow-roll models.
Note that the theory predicts that the
\textit{primordial} distribution of perturbations is Gaussian.
When the Universe becomes matter dominated, and density
perturbations undergo gravitational amplification, an initially
Gaussian distribution will become non-Gaussian \cite{Pee80}.
Such departures from initial Gaussianity have a specific
form and may be probed as consistency checks of inflation with
galaxy surveys that probe the matter distribution today.
\subsection{Brief Overview of Models}
A huge literature is devoted to construction of
inflationary models (for a comprehensive review, see
\cite{LytRio99}).
Here we follow the classification of Dodelson
et al \cite{DodKinKol97}. Models can be regarded as either large-field,
small-field, or hybrid models. Linear models live
at the border of large- and small-field models. In large-field
(small-field) models, the inflaton moves a distance $\Delta\phi$ that is
large (small) compared with the Planck mass during inflation.
Hybrid models introduce a second scalar field and allow a
broader range of phenomenology.
\begin{figure}
\centerline{\psfig{file=regions.ps,width=32pc}}
\caption{Regions in the $n_s$-$r$ plane occupied by the various
classes of inflationary models. (From References
\cite{DodKinKol97,Kin98}; their $n$ is our $n_s$.)}
\vskip-12pt
\label{fig:regions}
\end{figure}
The models can be distinguished experimentally by the values of
$V$, $\epsilon$, and $\eta$ that they predict, or equivalently by
the set of $V$, $r$, and $n_s$, as shown in
Figure \ref{fig:regions}.
Examples of large-field models are single-field models with
polynomial potentials, $V(\phi) \propto (\phi/\mu)^p$ (with
$p>1$), or in the $p\rightarrow \infty$ limit, exponential potentials, $V(\phi)
\propto \exp(\phi/\mu)$.\footnote{Exponential potentials are
sometimes referred to as ``power-law
inflation,'' since the scale factor grows as a power law during
inflation in these models.} The potentials in these models
resemble qualitatively the potential shown in Figure \ref{fig:potentials}\textit{a}.
These models have $V''>0$ and predict $\epsilon=[p/(p-2)]\eta>0$ and
$r\simeq 7 [p/(p+2)](1-n_s)$. Thus, a large tensor amplitude is
expected for large $p$ (and therefore for exponential potentials
as well) and for a sufficiently large deviation of $n_s$ from
unity.
Figure \ref{fig:potentials}\textit{b} shows a potential typical of a
small-field model. These are the types of potentials that often
occur in spontaneous symmetry breaking and can be approximated
by $V(\phi) \propto [1-(\phi/\mu)^p]$. These models have $V''<0$.
Demanding that the field
move a distance that is small compared with $m_{\rm Pl}$ requires that
$(\phi/\mu) \ll 1$, and in this limit,
$\epsilon=[p/2(p-1)]|\eta| (\phi/\mu)^p \ll \eta$, $\eta<0$, and
$r\simeq7(1-n_s) \epsilon/|\eta|$. Note that the slow-roll condition
$\phi \ll \mu$ implies that $\epsilon \ll 1$, so $\epsilon \ll
\eta$. It thus follows that $n_s \simeq 1+2\eta$, and that the
tensor amplitude in these models is expected to be very
small. Note that both small- and large-field models predict
$n_s<1$.
Linear models live at the border of small- and large-field
models. They have potentials $V(\phi) \propto \phi$ (i.e.\ they
have $V''=0$) and predict $\epsilon\simeq-\eta>0$ and
$r\simeq(7/3)(1-n_s)$.
Although hybrid models generally involve multiple scalar fields,
they can be parameterized by a single-field model with a
potential $V\propto [1+(\phi/\mu)^p]$. These models have
$\epsilon>0$ and
\begin{equation}
{\eta \over \epsilon} = {2 (p-1) \over p} \left({\phi
\over\mu} \right)^{-p} \left[ 1 + {p-2 \over 2(p-1)} \left(
{\phi \over \mu}\right)^p \right] >0.
\label{eq:hybrid}
\end{equation}
Unlike small- or large-field models, hybrid models can (although
are not required to) produce blue spectra, $n_s>1$. Although
both $r$ and $n_t$ depend only on $\epsilon$ and are thus
related, there is no general relation between $r$ and $n_s$ in
hybrid models. The tensor amplitude is only constrained to be
smaller than it is in exponential models.
\section{Cosmic Microwave Background Tests of Inflation}
\label{sec:cmbtests}
Photons from overdense regions
at the surface of last scatter are redder
since they must
climb out of deeper potential wells (the Sachs-Wolfe effect
\cite{SacWol67}). However, this is
really only one of a number of physical mechanisms that give
rise to temperature perturbations. We have also
mentioned that if primordial perturbations are adiabatic, then
the gas in deeper potential wells is hotter, and this partially
offsets the reddening due to the depth of the potential.
Density perturbations
induce peculiar velocities, and
these also produce temperature perturbations via Doppler shifts.
Growth of the
gravitational potential near the CMB surface of last scatter can
produce temperature anisotropies \cite{HuSug94a} [the early-time
integrated Sachs-Wolfe (ISW) effect], and so can the
growth of the gravitational potential at late times in a flat
cosmological-constant \cite{KofSta86} or open \cite{KamSpe94}
Universe (the late-time ISW effect).
Modern calculations of the CMB power spectra (the $C_l$) take
into account all of these effects. The cosmological
perturbation theory underlying these calculations has been reviewed
\cite{Bar80,KodSas84,MukFelBra92}, and solution of the
Boltzmann equations for the observed angular distribution of CMB
photons is discussed elsewhere \cite{MaBer95,Huetal95,SelZal96}.
Such calculations for the CMB power spectra from density
perturbations were developed in a series of pioneering papers
{}from 1970 until the late 1980s
\cite{PeeYu70,WilSil80,WilSil81,SilWil81,BonEfs84,VitSil84,
BonEfs87,Hol89},
and these calculations have been refined extensively in the post-\COBE\
era. Similar calculations can also be carried out for the CMB
power spectra from gravitational waves
\cite{AbbWis84,MilVal86,HarZal93,CriDavSte93,FrePolCol94,NgNg95,Kos96,
KamKosSte97b,ZalSel97,HuWhi97}.
The calculations for both scalar and tensor power spectra
require solution of a series of several
thousand coupled differential equations for the perturbations to
the spacetime metric, densities and velocities of baryons and
cold dark matter, and the moments of the
CMB photon and neutrino distributions. A code for carrying out
these calculations (that required several hours for each model)
was made publicly available \cite{Ber95}. Hu \& Sugiyama
\cite{HuSug94} came up with useful semianalytic fits to the
numerical calculations that provided some physical intuition
into the numerical results. More recently, Seljak \& Zaldarriaga
\cite{SelZal96,SelZal97} developed a line-of-sight
approach that speeded up the numerical calculations by several orders of
magnitude. A code (CMBFAST) was made publicly available and has
become widely used.
\begin{figure}
\centerline{\psfig{file=colormod.ps,width=32pc}}
\caption{Theoretical predictions for cosmic-microwave-background
temperature angular power spectra as a function
of multipole moment $l$ for models with primordial
adiabatic perturbations.
Each graph shows the effect
of variation of one of these parameters. In the lower
right panel,
$\Omega\equiv\Omega_0+\Omega_\Lambda=1$. (From
Reference \cite{Junetal96b}.)}
\vskip-12pt
\label{fig:models}
\end{figure}
Given the values of the classical cosmological parameters (e.g.\
the nonrelativistic matter density $\Omega_0$, cosmological
constant $\Omega_\Lambda$, and baryon density $\Omega_b$, all in
units of the critical density, and the Hubble parameter $h$ in units of
100~km~sec$^{-1}$~Mpc$^{-1}$), and primordial scalar and tensor
power spectra, $P_s(k)$ and $P_t(k)$, it is straightforward to
calculate the $C_l$ with the machinery described above.
Figure~\ref{fig:models} shows results of such calculations for
models with a Peebles-Harrison-Zeldovich (i.e.\ $n_s=1$) power
spectrum of primordial adiabatic perturbations. Each panel
shows the effect of independent variation of one of the
cosmological parameters. As illustrated, the height, width, and
spacing of the acoustic peaks in the angular spectrum depend on
these (and other) cosmological parameters.
The
wiggles\footnote{These are sometimes referred to inaccurately
as ``Doppler'' peaks, but are more accurately referred
to as acoustic peaks. They are sometimes called Sakharov
oscillations in honor of the scientist who first postulated the
existence of photon-baryon oscillations in the primordial plasma
\cite{Sak65}. The existence of these peaks in the CMB power
spectrum was, to the best of our knowledge, first identified by
Sunyaev \& Zeldovich \cite{SunZel70} and Peebles \& Yu
\cite{PeeYu70}.}
come from oscillations in the photon-baryon fluid at
the surface of last scatter. Consider an individual Fourier
mode of an initial adiabatic density perturbation.
Because the density perturbation is nonzero initially, this
mode begins at its maximum amplitude. The amplitude remains
fixed initially when the wavelength of the mode is larger than
the Hubble radius. When the Universe has expanded enough that
the Hubble radius becomes larger than the wavelength of this
particular mode, then causal physics can act, and the amplitude
of this Fourier mode begins to oscillate as a standing acoustic
wave \cite{Sak65}. Since modes with smaller wavelengths come
within the horizon earlier and oscillate more rapidly,
they have at any given time undergone more oscillations
than longer-wavelength modes have. The peaks evident in
Figure \ref{fig:models} arise because modes of different
wavelength are at different points of their oscillation cycles
\cite{SunZel70}. The first peak corresponds to the mode that has had
just enough time to come within the horizon and compress
once. The second peak corresponds to the mode that is at its maximum
amplitude after the first compression, and so forth.
\subsection{Determination of the Geometry}
These acoustic peaks in the CMB temperature power spectrum can
be used to determine the
geometry of the Universe \cite{KamSpeSug94}. The angle
subtended by the horizon at the surface of last scatter is
$\theta_H \sim \Omega^{1/2} \;1^\ifmmode ^\circ\, \else $^\circ\,$ \fi$, where
$\Omega=\Omega_0+\Omega_\Lambda$ is the total density (objects
appear to be larger in a closed Universe than they would in a
flat Universe, and smaller in an open Universe than they would
in an flat Universe).
Moreover, the peaks in the CMB
spectrum are due to causal processes at the surface of last
scatter. Therefore, the angles (or values of $l$) at which the
peaks occur determine the geometry of the Universe. This is
illustrated in the top left panel of Figure~\ref{fig:models}, where the CMB spectra
for several values of $\Omega$ are shown. As illustrated in the
other panels, the angular position of the first peak is
relatively insensitive to the values of other undetermined (or
still imprecisely determined) cosmological parameters such as
the baryon density, the Hubble constant, and the cosmological
constant.
Small changes to the spectral index $n_s$ tilt the entire
spectrum slightly to smaller (larger) $l$ for $n_s<1$ ($n_s>1$),
and the location of the first peak is only weakly affected.
Gravitational waves would only
add to the temperature power spectrum at $l\ll 200$ (as
discussed below in Section 4.4). Therefore,
although gravitational waves could affect the height of the
peaks relative to the normalization at small $l$, the locations
would not be affected. It is plausible that an early generation
of star formation released a sufficient flux of ionizing
radiation to at least partially reionize the Universe, and if
so, these ionized electrons would rescatter some fraction
$\tau$ of the CMB photons. A variety of theoretical arguments
suggest that a fraction $\tau = {\cal O}(0.1)$ of CMB photons
were rescattered \cite{KamSpeSug94,TegSilBla94,HaiLoe97} (and
the amplitude of anisotropy at degree scales observed already
supports this). Although
reionization would damp the peaks by a factor $e^{-2\tau}$, as
indicated by the curve labeled ``reion'' in the top left
panel of Figure \ref{fig:models} (but note that the figure assumes
$\tau=1$), the location of the peaks would remain unchanged.
Therefore, if peak structure is observed in the CMB power
spectrum, determination of the location of the first peak will
provide a robust determination of the geometry of the Universe
\cite{KamSpeSug94}.
\subsubsection{Open inflation}
The most distinctive signature of an open-inflation model would be a
low-$\Omega_0$ CMB power spectrum from adiabatic perturbations such as
one of those shown in the top left panel of Figure \ref{fig:models},
in which the first peak is shifted to larger $l$.
Open inflation would also produce an increase in large-angle
anisotropy from the integrated Sachs-Wolfe effect
\cite{LytSte90,KamSpe94}, but different open-inflation
models make different predictions about the large-angle anisotropy.
Moreover, determination of the CMB power spectrum at these large
angular scales is cosmic-variance limited, so it is unlikely
that large-angle CMB anisotropies alone will be able to provide
a robust test of open-inflation models. The ISW effect may
alternatively be identified by cross-correlation of the CMB with
some tracer of the mass density along the line of sight
\cite{CriTur96,Kam96}, such as the X-ray background
\cite{BouCriTur98,KinKam98} or possibly weak-lensing maps
\cite{ZalSel99} (as discussed further in Section
\ref{sec:cosmologicalconstant}).
\begin{figure}
\centerline{\psfig{file=huwhite.ps,width=32pc}}
\caption{The angular power spectrum for an inflationary model with
primordial adiabatic perturbations and for another with
primordial isocurvature perturbations.
\textit{Solid line}, cold dark matter and inflation;
\textit{dashed line}, axion isocurvature. (From
Reference \cite{HuWhi96a}.)}
\vskip-12pt
\label{fig:huwhite}
\end{figure}
\subsection{Adiabatic Versus Isocurvature Modes}
The physics described above yields a distinctive pattern in the peak
structure of the CMB power spectrum, and this leads to an important
test of inflation. If primordial perturbations are
isocurvature rather than adiabatic,
then when a given Fourier mode comes within the
horizon and begins to oscillate, it begins to oscillate from its
minimum (rather than maximum) amplitude. Thus, the phase of its
oscillation differs by $\pi/2$ from what it would be if the
perturbation were adiabatic. As a result, the locations of the peaks in
the CMB power spectrum in isocurvature models differ in phase from
what they would be in adiabatic models \cite{HuSug95,HuWhi96a,Kos98}, as shown
in Figure \ref{fig:huwhite}.
It has also been shown that the
relative locations of the higher peaks differ in adiabatic and
isocurvature models, independent of the shift in the locations of the
peaks due to the geometry \cite{HuWhi96}.
\subsubsection{Axion inflation}
When the matter power spectrum is normalized to
the amplitude of galaxy clustering, isocurvature models with
nearly scale-invariant primordial power spectra (e.g.\ axion
isocurvature or ``axion inflation'' models) produce roughly six
times the CMB anisotropy seen by \COBE\ \cite{KodSas86,EfsBon87}
(since there is no cancellation between the effects of the
intrinsic temperature and the potential-well depth at the
surface of last scatter) and are thus ruled out.
\subsection{Coherent Perturbations and Polarization}
Each Fourier component of the density field induces a Fourier
component of the peculiar-velocity field, and the oscillations
of these peculiar velocities are out of phase with the
oscillations in the density perturbation (just as the velocity
and position of a harmonic oscillator are out of phase). These
peculiar velocities induce temperature anisotropies (via the
Doppler effect) that are thus out of phase with those from
density perturbations. This Doppler effect fills in the troughs
in the $C_l^{\rm TT}$, which would otherwise fall to zero.
The CMB polarization is related to
the peculiar velocity at the surface of last scatter
\cite{ZalHar95}, so the peaks in the polarization power
spectrum (from density perturbations) are out of phase from those in
the temperature power spectrum and fall close to zero
(Figure \ref{fig:clsplot}). This relative
positioning of the temperature and polarization peaks is a
signature of coherent perturbations (rather than those
produced, for example, by the action of topological defects, as
discussed below) \cite{Kos98}.
Zaldarriaga \& Spergel \cite{SpeZal97} argue that the location of the
first peak in the polarization power spectrum provides a test of
primordial perturbations as the origin of structure and thus of
inflation. If some causal mechanism (such as topological defects)
produced large-scale structure, the the first peak would have to occur
at smaller angular scales in order to be within the horizon at the
surface of last scatter (see below). A peak so close to the causal
horizon could only occur with super-horizon-sized
primordial perturbations, for which inflation is the only causal
mechanism.
\begin{figure}[htbp]
\centerline{\psfig{file=clt0.ps,width=7in}}
\bigskip
\caption{
Theoretical predictions for the four nonzero cosmic-microwave-background
temperature-polarization spectra as a function
of multipole moment $l$. \textit{Solid curves} are the
predictions for a \COBE-normalized inflationary model
with no reionization and no gravitational waves for
$h=0.65$, $\Omega_b
h^2=0.024$, and $\Lambda=0$. \textit{Dotted curves} are the
predictions that would be obtained if the \COBE\
anisotropy were due entirely to a stochastic
gravity-wave background with a flat scale-invariant
spectrum (with the same cosmological parameters).
The panel for $C_l^{\rm CC}$
contains no dotted curve because scalar perturbations
produce no ``C'' polarization component; instead,
the \textit{dashed line} in the \textit{bottom right panel} shows a
reionized model with optical depth $\tau=0.1$ to the
surface of last scatter. (From Reference \cite{KamKos98}.)
}
\label{fig:clsplot}
\end{figure}
\subsection{Polarization and Gravitational Waves}
Gravitational waves are usually detected by observation of the motion
they induce in test masses. The photon-baryon fluid at the
surface of last scatter acts as a set of test masses for
detection of gravitational waves with wavelengths comparable to
the horizon, such as those predicted by inflation. These
motions are imprinted onto the temperature and polarization of
the CMB. The top left panel of Figure \ref{fig:clsplot}
(\textit{solid curve}) shows the
temperature power spectrum for a \COBE-normalized flat
scale-invariant ($n_t=0$) spectrum of gravitational waves. It
is flat and relatively featureless for $l\la70$.
The dropoff at $l \ga70$ is due to the fact that the amplitudes
of gravitational-wave modes that enter the horizon before the
epoch of last scatter have decayed with the expansion of the
Universe. Unfortunately, cosmic variance from scalar
perturbations provides a fundamental limit to the sensitivity of
CMB temperature maps to tensor perturbations \cite{KnoTur93}.
Even if all other cosmological parameters are somehow fixed,
a perfect temperature map can never detect an
inflaton-potential height smaller than one-tenth the upper limit
provided by \COBE\ \cite{Junetal96b}. More realistically, the
effects of gravitational waves and reionization on the
temperature power spectrum are similar and difficult to
disentangle, so improvements to the current \COBE\ sensitivity
to gravitational waves is unlikely with a temperature map alone.
However, with a polarization map of the CMB, the scalar and
tensor contributions to CMB polarization can be geometrically
decomposed in a model-independent fashion, and the
cosmic-variance limit present in temperature maps can thereby be
circumvented \cite{Ste96,KamKosSte97a,SelZal97}. Scalar perturbations
have no handedness, so they cannot give rise to a curl
component. On the other hand, tensor perturbations do
have a handedness, so they induce a curl component. Therefore,
if any curl coefficient, $a_{(lm)}^{\rm C}$, is found to be
nonzero, it suggests the presence of gravitational
waves.\footnote{A curl component
may also be due to vector perturbations.
Although topological-defect models may excite such modes, they
do not arise in inflationary models.}
To illustrate, Figure~\ref{fig:clsplot} shows the four nonzero
temperature-polarization power spectra. The \textit{dotted curves}
correspond to a \COBE-normalized inflationary model
with no gravitational waves. The \textit{solid
curves} show the spectra for a \COBE-normalized stochastic
gravitational-wave background.
\subsubsection{Detectability of gravitational waves: curl
component only}
Consider a mapping experiment that measures the
polarization on the entire sky with a temperature sensitivity
$s$ (which has units $\mu$K~$\sqrt{\rm sec}$) for a time $t_{\rm
yr}$ years. If only the curl component of the polarization is
used to detect tensor perturbations, then such an experiment can
distinguish a tensor signal from a null result at the $2\sigma$
level if the inflaton potential height is \cite{KamKos98}
\begin{equation}
V \mathrel{\mathpalette\fun >} (4\times 10^{15}\, {\rm GeV})^4 \, t_{\rm yr}^{-1} \,
(s/\mu{\rm K}\, \sqrt{\rm sec})^2.
\label{eq:tensordetectable}
\end{equation}
Equation~\ref{eq:tensordetectable} indicates that to access an
inflaton-potential height not already excluded by \COBE\
requires a detector
sensitivity $s\la35\,t_{\rm yr}^{1/2}\,\mu$K$\sqrt{\rm
sec}$. To compare this with realistic values, the effective
sensitivity of MAP is $s\simeq150\,t_{\rm
yr}^{1/2}\,\mu$K$\sqrt{\rm sec}$ and that for Planck is about
$s\simeq35\,t_{\rm yr}^{1/2}\,\mu$K$\sqrt{\rm sec}$, and
technological developments have improved the detector
sensitivity roughly an order of magnitude per decade for the
past several decades. Even better sensitivities may be possible
with deep integration on a smaller region of the sky.
\subsubsection{Reionization}
In some sense, Equation~\ref{eq:tensordetectable} is conservative
because even a
small amount of reionization will significantly increase the
polarization signal at low $l$ (indicated by the \textit{dashed curve}
in the CC panel of Figure~\ref{fig:clsplot} \cite{Zal97}).
If $\tau=0.1$, then the sensitivity to tensor modes is increased
by roughly a factor of five \cite{KamKos98}.
\subsubsection{Full polarization and temperature spectra}
Although searching only for the curl component provides a
model-independent probe of tensor modes, a
stochastic gravitational-wave background leads to specific
predictions for all four nonzero temperature-polarization power
spectra (Figure \ref{fig:clsplot}). Fitting an
inflationary model to the entire set of temperature and
polarization power spectra can improve tensor detectability,
especially at poorer sensitivities, albeit with the introduction
of some model dependence. For detector sensitivities $s\mathrel{\mathpalette\fun >}
15\,t_{\rm yr}^{1/2}\,\mu$K$\sqrt{\rm sec}$, the sensitivity to
a tensor signal is improved by factors of a few or so
\cite{KamKos98}, depending on the cosmological model, whereas
for detector sensitivities $s\mathrel{\mathpalette\fun <} 15\,t_{\rm
yr}^{1/2}\,\mu$K$\sqrt{\rm sec}$, the sensitivity is attributable
almost
entirely to the CC power spectrum and approaches the limit in
Equation~\ref{eq:tensordetectable}.
\begin{figure}[htbp]
\centerline{\psfig{file=newmodel.ps,width=5.5in}}
\bigskip
\caption{Simulated $2\sigma$ error ellipses that would be
obtained by a cosmic-variance-limited temperature map,
the Planck Surveyor (with polarization), and an
experiment with three times the sensitivity of Planck.
This assumes an inflationary model with $r=0.01$
and $n_s=0.95$ and an optical depth to the surface of
last scatter of $\tau=0.05$. \textit{Shaded regions}
indicate the
predictions of various inflationary models. \textit{Solid
horizontal curve} indicates the regions of this
logarithmic parameter space that would be accessible
with a putative polarization experiment with 30 times
the Planck instrumental sensitivity \cite{KamKos98}.
Even better sensitivities may be possible with deep
integration on a smaller region of the sky.
(From Reference \cite{Kin98}.)
}
\label{fig:Plancklog}
\end{figure}
\subsubsection{Measurement of inflationary observables}
Several authors have addressed the question of
how precisely the inflationary observables can be reconstructed
in the case a positive detection of the stochastic
gravitational-wave background with only a temperature map
\cite{Kno95,Junetal96b,BonEfsTeg97,DodKinKol97} and with a
polarization map as well \cite{ZalSelSpe97,Lidetal97b,Kin98}.
We follow the discussion of Ref. \cite{Kin98}.
Figure \ref{fig:Plancklog} shows the $2\sigma$ error ellipses
that would be obtained by the Planck Surveyor using the
temperature only (i.e. the cosmic-variance limit) and with the
polarization, assuming a gravitational-wave background with
$r\simeq0.01$ and $n_s\simeq0.9$. (A larger gravitational-wave
amplitude would be detectable, as shown in Figures~3--6 in
Reference \cite{Kin98}.) The $2\sigma$ cosmic-variance limit from a
temperature map is shown as is the $2\sigma$ constraint to the
parameter space expected for Planck (with polarization).
Although such a tensor
amplitude cannot be distinguished from a null result, the figure
shows (the dark shaded region) that a hypothetical experiment with three times
the Planck polarization sensitivity could discriminate between a
such a tensor signal and a null result. It would also
discriminate between a single small-field model and a hybrid
model. Of course, the sensitivity to tensor modes can be
improved as the instrumental sensitivity is improved, as
indicated by Equation~\ref{eq:tensordetectable}. For example, the thin
horizontal line at $r=0.001$ shows the smallest value of $r$
that could be distinguished from a null result by a hypothetical
one-year experiment with an instrumental sensitivity
$s=\mu$K$\sqrt{\rm sec}$, roughly 30 times that of Planck
\cite{KamKos98}. A
null result from such an experiment would suggest that if
inflation occurred, it would have required a small-field model.
\subsection{Gaussianity}
The prediction of primordial Gaussianity or of some specific
small deviations from Gaussianity can be probed with the CMB
angular bispectrum or higher $n$-point correlation functions
discussed above. A nonzero large-angle CMB
bispectrum may arise from the integrated Sachs-Wolfe effect if
there is a cosmological constant \cite{Ganetal94}. Such a
bispectrum, as well as that probed by the Sunyaev-Zeldovich
effect, may be discernible via cross-correlation between
gravitational-lensing maps and the CMB \cite{GolSpe98}. A more
powerful test of inflation models may arise from probing
the bispectrum induced by nonlinear evolution
at the surface of last scattering (S Winitzki, A Kosowsky, DN Spergel,
manuscript in preparation).
Ferreira et al and Pando et al \cite{FerMagGor98,PanValFan98}
recently claim to have already found some signature of
non-Gaussianity in the \COBE\ maps. In particular,
Ferreira et al \cite{FerMagGor98} find $I_l^3 \sim1$ for
$l\sim16$ (although it is still not clear if the effect is real
\cite{KamJaf98}). If this result is correct, then the simplest
slow-roll inflation models are not viable (see
Equation~\ref{eq:Il3prediction}).
\section{Topological-Defect Models}
\label{sec:defects}
The leading alternative
to structure-formation models based on inflation have been
those based on topological defects, particularly
cosmic strings \cite{Zel80,Vil81,SilVil84,Tur85,HinKib95},
global monopoles \cite{BarVil89,BenRhi91},
domain walls \cite{HilSchFry89,PreRydSpe89},
and textures \cite{Tur89,Tur91}
(for reviews, see \cite{Vil85,VilShe94}).
Defect models postulate a phase transition
in the early Universe that leads to a vacuum manifold with
nontrivial topology; the type of defect depends
on the specific topology (see \cite{HinKib95} for a
review). Since defect formation is a process
governed by causal physics, the vacuum state of the field
must be uncorrelated on scales larger than the horizon at
the time of the phase transition, guaranteeing the
formation of defects with a characteristic length scale
of the horizon (the ``Kibble mechanism'' \cite{Kib76}).
The simplest defects are domain walls, which arise in theories
with a discrete symmetry. Domain-wall models are not
viable because their energy densities are
large enough to produce larger CMB temperature fluctuations
than those observed \cite{ZelKozOku74,SteTur89,TurWatWid91}.
Cosmic strings are stable defects that arise in gauge models
with a $U(1)$ symmetry. They
produce density perturbations by their gravitational
interactions with ordinary matter.
Global-monopole and texture models are
unstable defects that arise in models with a perfect
global symmetry. They provide two mechanisms for structure
formation: (\textit{a}) the energy-density provided by misalignment of
scalar fields as causally disconnected regions of the Universe
come into causal contact, and (\textit{b}) the explosive events that
occur when the topological defects unwind. Non-topological
texture models \cite{Vil82,Pre80,TurSpe91,Jaf94} postulate an
even higher global symmetry and seed structure via scalar-field
alignment even though no topological defects are formed.
Generically, one expects quantum gravity to violate global
symmetries to the level that would render global-monopole,
texture, and scalar-field-alignment models unworkable
\cite{KamMar92,Holetal92}. If it could be shown that such
models do seed large-scale structure, valuable information on
Planck-scale physics would thus be provided.
\subsection{Cosmic-Microwave-Background Power Spectra in Defect Models}
In contrast to inflationary models, which lay down an initial
spectrum of density perturbations, defect models
produce perturbations actively throughout the history of the
Universe. This generally leads to loss of coherence in the
perturbations and a corresponding smoothing of the acoustic peaks
\cite{Albetal96,Magetal96a,HuWhi97c}. Defect-model perturbations are also
causal, being generated by physical processes
inside the horizon
\cite{SteVee90,Tur96b,Tur96c,HuSpeWhi97,DurKun98}.
And finally,
primordial perturbations in defect models more closely resemble
those in primordial-isocurvature rather than those in adiabatic models
\cite{SteVee90,DurSak97,HuSpeWhi97}.
Moreover, the action of topological defects
generically produces vector and tensor perturbations which
increase the anisotropy on small angular scales
\cite{PenSelTur97,Alletal97}, further suppressing any peak
structure (although it may produce some characteristic C
polarization \cite{SelPenTur97}).
\begin{figure}[htbp]
\centerline{\psfig{file=defects.ps,width=5.5in}}
\bigskip
\caption{Cosmic-microwave-background power spectra from topological-defect models.
\textit{Solid line}, total; \textit{dotted}, scalar;
\textit{short-dashed}, vector; \textit{long-dashed}, tensor.
(From Reference \cite{SelPenTur97}.)}
\label{fig:defects}
\end{figure}
Until recently, different groups obtained
different results about the extent to which acoustic peaks
exist in defect models, and under what circumstances
\cite{CriTur95,Tur96b,DurGanSak96,DurZho96,Tur96b,Tur96c}.
Calculations of CMB power spectra based on large simulations of
a variety of defect sources have now been performed
\cite{Alletal97,PenSelTur97,ConHinMag99} (see
Figure \ref{fig:defects} for some). The numerical results
indicate that the acoustic peaks are washed out.
At this point, it appears that the simplest defect models are
inconsistent with the observed CMB fluctuations and the large-scale
structure traced by galaxy surveys. Although this could have
been inferred from the generic arguments discussed in the
Introduction \cite{JafSteFri94}, it has been supported by these
more recent precise calculations
\cite{AlbBatRob97,Alletal97,PenSelTur97}. The question now is
whether any more complicated (or ``sophisticated'') defect models be
viable.
Albrecht et al \cite{AlbBatRob98a,AlbBatRob98b} have
suggested that a cosmological constant might help improve
concordance with current data. However, suppose the CMB
temperature power spectrum continues to look increasingly like that
caused by inflation (i.e.\ with identifiable acoustic peaks), as
new data seem to suggest.
If so, can any defect model reproduce such a power spectrum?
Turok \cite{Tur96c} produced
a power spectrum with a phenomenological defect model that
closely mimicked an inflation power spectrum, and Hu
\cite{Hu99} has invented a similar isocurvature model.
But it is hard to see how
to position the acoustic peaks in isocurvature-like models at
the same angular scales as in adiabatic models without some
rather artificial initial conditions
\cite{HuSpeWhi97,Magetal96b}.
It is also difficult to simultaneously account for the fluctuation
amplitude in the CMB and galaxy surveys, unless there is a
breaking of scale invariance \cite{Pen98} (possibly from some finite breaking
of the global symmetry \cite{KamMar92,Holetal92}).
It may, in fact, be possible to construct some causal
models that produce peaks in the CMB power spectrum
\cite{DurKun98,DurSak97}, but it is unclear whether
fluctuations that mimic a specific inflationary model can be
produced, particularly when additional constraints from
polarization are taken into account.
Finally, it should be noted that hybrid models with both
primordial adiabatic perturbations and defects have been
entertained \cite{Jen96,LinRio97,AveCalMar98}.
\subsection{Non-Gaussianity}
Topological-defect models may also be distinguished
by the non-Gaussian signatures they produce in the CMB.
Because the evolution of topological defects is
nonlinear, they generically produce non-Gaussian structures in
the CMB. Put another way, the production of defects via the
Kibble mechanism is a Poisson process; the number of defects
within any volume in the Universe is Poisson distributed. The
central-limit theorem guarantees that as the number density of
defects becomes large, the distribution should become
increasingly Gaussian. Thus cosmic-string models should look
more like Gaussian perturbations than textures should, since the Kibble
mechanism produces roughly one texture per 25 Hubble volumes as
opposed to roughly one cosmic string per Hubble volume
\cite{SchBer91,Per93b,BenRhi93,GilPer95,SchSch95}.
In the large-$N$
limit of the $O(N)$ sigma model, the clearest signature of
non-Gaussianity from scalar-field alignment is at large angular
scales \cite{Jaf94}; on small distance scales, the theory looks roughly
Gaussian. Constraints to non-Gaussianity from the
galaxy distribution have already posed problems for
scalar-field-alignment models for several years.
Since defects are
coherent structures, they can produce
corresponding coherent structures
in the CMB temperature anisotropy. For example, a cosmic string
can produce a linear discontinuity in the CMB temperature
\cite{KaiSte84}, which can be searched for most efficiently
through statistics tailored to match this particular signal
\cite{MoePerBra94,Per97,Per98}. Textures might form large hot
spots \cite{TurSpe90,Tur96a}.
\section{Dark Matter}
\label{sec:darkmatter}
The CMB can potentially provide a wealth of information about the dark
matter known to dominate the mass of the Universe.
The smallness of the amplitude of CMB temperature
fluctuations has for a long time provided some of the strongest evidence
for the existence of dark matter. In a low-density Universe, density
perturbations grow when the Universe becomes matter-dominated and end
when it becomes curvature-dominated. If the luminous matter
($\Omega_{\rm lum} \sim10^{-3}$) were all the mass in the Universe,
then the epoch of structure formation would be too short to allow
density perturbations to grow from their early-Universe amplitude,
fixed by \COBE, to the amplitude observed today in galaxy
surveys.
More precise measurements of the CMB power spectrum hold the
promise of providing much more detailed information about the
properties and distribution of dark matter.
There are currently several very plausible dark-matter
candidates that arise from new particle physics, and
some evidence has already been claimed for the
existence of several of these. For example, some observational
evidence points to the existence of a cosmological constant
\cite{Peretal97,Peretal99,Rieetal98}, and the LSND experiment
suggests that massive neutrinos may constitute a significant fraction
of the mass of the Universe \cite{Athetal95,Athetal96}.
Moreover, there are good arguments that a significant fraction
of the mass in galactic halos
is made of some type of
cold-dark-matter particle, e.g.\ weakly interacting massive
particles (WIMPs) \cite{JunKamGri96} or axions \cite{Tur90,Raf90}.
\subsection{Cold Dark Matter}
A number
of dynamical measurements suggest that the nonrelativistic-matter
density is $\Omega_0\ga0.1$, whereas big-bang nucleosynthesis suggests a
baryon density of $\Omega_b\la0.1$. Observations of
X-ray emission from galaxy clusters suggest that the
nonrelativistic matter in clusters outweighs the baryonic matter
by a factor of three or more \cite{Whietal93}, and weak lensing
of background galaxies by clusters directly reveals
large amounts of dark matter \cite{TysKocDel98}. This evidence
strongly indicates the existence of some nonbaryonic dark
matter. By fitting the power spectra from MAP and Planck to
theoretical predictions, one should simultaneously be able to
determine both $\Omega_0 h^2$ and $\Omega_b h^2$ to far better
precision than that obtained by current observations
\cite{Junetal96b,BonEfsTeg97,ZalSelSpe97}. If a substantial
fraction of the mass in the Universe is in fact made of
nonbaryonic dark matter (e.g.\ WIMPS or axions), then
it will become evident after MAP and Planck. Unfortunately,
there is no way to discriminate between WIMPs and axions with
the CMB.
\subsection{Neutrinos}
One of the primary goals of experimental particle
physics is pursuit of a nonzero neutrino mass. Some
recent (still controversial) experimental results suggest that
one of the neutrinos may have a mass of ${\cal O}(5\,{\rm eV})$
\cite{Athetal95,Athetal96}. There have been some arguments
(again, still
controversial) that such a neutrino mass is exactly what
is required to explain apparent discrepancies between
large-scale-structure observations and the simplest
inflation-inspired standard-CDM model
\cite{ShaSte84,DalSch92,DavSumSch92,Klyetal93,Prietal95,BonPie98}.
If the neutrino does indeed have a mass of ${\cal O}(5 \, {\rm
eV})$, then roughly 30\% of the mass in the Universe is
in the form of light neutrinos. These neutrinos will affect the
growth of gravitational-potential wells near the epoch of last
scatter, thus leaving an imprint on the CMB angular
power spectrum \cite{DodGatSte96,MaBer95,Lopetal98a}. The
effect of a light neutrino
on the power spectrum is small, so other cosmological parameters
that might affect the shape of the power spectrum at larger
$l$ must be known well. Eisenstein et al \cite{EisHuTeg99} argue
that by combining measurements of the CMB power spectrum with
those of the mass power spectrum measured by, for instance, the Sloan
Digital Sky Survey, a neutrino mass of ${\cal O}(5 \, {\rm eV})$
can be identified. The CMB may
constrain the number of noninteracting relativistic
degrees of freedom in the early Universe \cite{Junetal96b}.
Although weaker than
the bound from big-bang nucleosynthesis \cite{SteSchGun77,
CopSchTur97}, the CMB
probes a different epoch ($T\sim$eV rather than $T\sim$MeV) and
may thus be viewed as complementary.
\subsection{Cosmological Constant}
\label{sec:cosmologicalconstant}
Some recent evidence seems to point to the existence of an
accelerating expansion, possibly due to a nonzero cosmological
constant (\cite{Peretal97,Peretal99}; for a review
of the cosmological constant, see \cite{CarPreTur92}).
The CMB may help probe the existence of a cosmological constant
in a number of ways. As discussed
above, if adiabatic perturbations are responsible for
large-scale structure, then the position of the first acoustic
peak in the CMB power spectrum
provides a model-independent probe of the total density,
$\Omega=\Omega_0 + \Omega_\Lambda$ \cite{KamSpeSug94}. In
contrast, the supernova measurements of the Hubble diagram at
large redshifts determine primarily the deceleration
parameter $q_0 = \Omega_0/2 - \Omega_\Lambda$, so the two
measurements together can give tight limits on both $\Omega_0$
and $\Omega_\Lambda$ individually
\cite{Efsetal98,Garetal98,Teg98,TegEisHu98,Whi98,Lin98}.
As the bottom panels of Figure \ref{fig:models} show, variations to
$\Omega_0$ and $h$ affect the the height and width of the first
acoustic peak; the dependence is more precisely on
the quantity $\Omega_0 h^2$. Thus, if the Hubble constant is
known, then the CMB can determine $\Omega_0$ and $\Omega$ (from
the peak location) and therefore the cosmological constant
$\Omega_\Lambda$.
A cosmological constant may also be distinguished from the CMB
via the additional large-angle anisotropy it produces via the
ISW effect \cite{KofSta86} from density perturbations at
redshifts $z\mathrel{\mathpalette\fun <}$few. If there is a cosmological constant, there
should be a
cross-correlation between the CMB temperature and some tracer of
the mass distribution, e.g.\ the extragalactic X-ray background
\cite{BouCriTur98} or weak lensing \cite{ZalSel99}, at these
redshifts \cite{CriTur96} (the same also occurs in an open
Universe \cite{Kam96,KinKam98}). An experimental upper limit to
the amplitude of this cross-correlation \cite{BouCriTur98} can
already be used to constrain $\Omega_0$, with some assumptions
about the bias of sources that give rise to the extragalactic
X-ray background. If $\Omega_0\simeq0.3$ (either in an open or
a flat cosmological-constant model), then these X-ray sources
can be no more than weakly biased tracers of the mass
distribution \cite{KinKam98}.
\subsection{Rolling Scalar Fields}
The supernova evidence for an accelerating expansion has
engendered a burst of theoretical activity on exotic forms of
matter with an equation of state $p<-\rho/3$ (i.e.\ the equation of state needed
for $q_0<0$). The simplest possibility is of course a
cosmological constant. However, as explained in
Section \ref{sec:scalarfields}, a rolling scalar field
may also provide
such an equation of state, provided the scalar field is not
rolling too quickly. An almost endless variety of equations of
state---and expansion histories---are possible in principle,
given the freedom to choose the scalar-field potential and the
initial conditions. This idea is variously referred to in the literature
as rolling-scalar-field, variable-cosmological-constant,
x-matter, generalized-dark-matter,
loitering-Universe, and/or quintessence models
\cite{RatPee88,SahFelSte92,SugSat92,Frietal95,CobDodFri97,
SilWag97,TurWhi97,ChiSugNak97,CalDavSte98,ChiSugNak98}.
Additional work has explored attractor solutions based on
exponential potentials
\cite{LucMat85,Wet88,WanCopLid93,FerJoy97,CopLidWan98,LidSch99} or
``tracker-field'' solutions \cite{ZlaWanSte99,SteWanZla99} that
attempt to explain why the matter density would be comparable to
a scalar-field energy density today.
Because the expansion rate at decoupling in such models is the
same as in cosmological-constant models with the same
$\Omega_0$, the peak structure in the CMB is virtually
indistinguishable from that in cosmological-constant models
\cite{Hueetal99}. However, perturbations in the scalar field
track perturbations to the matter density on large scales in
such a way that the large-angle ISW effect that appears in
cosmological-constant models is canceled by the effect of
scalar-field perturbations \cite{CalDavSte98}.
Data from cosmological observations, particularly
supernova measurements of the expansion history and
measurements of the power spectrum through large
galaxy surveys, may in principle
be used to break these degeneracies
\cite{Huetal99,Wanetal99}.
\section{Other Constraints on Particle Physics}
\begin{figure}[htbp]
\centerline{\psfig{file=huspectr.ps,width=5.5in,angle={-90}}}
\bigskip
\caption{Constraints to the mass-lifetime plane for particles
decaying to photons from FIRAS constraints to
distortions to the CMB blackbody spectrum. \textit{Solid curve} is
the numerical result; \textit{dashed curves} show various
approximations. The quantity $n_X/n_\gamma$ is the
initial ratio of the particle number density to the
photon number density. (From Reference
\cite{HuSil93a}.)}
\label{fig:decaylimits}
\end{figure}
\subsection{Decaying Particles}
As discussed in Section \ref{sec:freqspectrum},
FIRAS limits to $\mu$ and $y$ distortions limit the injection
of energy into the early Universe and can thus be
used to constrain the mass-lifetime plane of particles that
decay to electromagnetically interacting particles (as shown in
Figure~\ref{fig:decaylimits}) \cite{HuSil93a,Elletal92}.
The CMB power spectrum can also constrain decaying particles.
For example, a neutrino of mass $\ga10$ eV that decays to
relativistic particles with a lifetime $\tau \simeq 10^{13-17}$
alters the expansion rate of the Universe between recombination
and today and thus produces large-angle anisotropy (via the ISW
effect) in disagreement with observations
\cite{Lopetal98b,Han98b,Han98c}.
\subsection{Time Variation of Fundamental Constants}
A number of ideas for new physics postulate that some of the
fundamental constants of nature, such as the fine-structure
constant $\alpha$, may actually be varying (for a review, see
\cite{VarPot95}). Such a variation
could be caused by the cosmological evolution
of compact spatial dimensions in string theory
or Kaluza-Klein theories \cite{Mar84,Bar87,DamPol94}
or through scalar fields coupled to electromagnetism \cite{Car98}.
Limits of $|\Delta\alpha/\alpha| \mathrel{\mathpalette\fun <} 10^{-7}$ were provided by
the natural nuclear reactor at Oklo \cite{Shy76,DamDys96},
and observations of atomic- and molecular-line positions at
high redshifts \cite{Sav56} provide limits of
$|\Delta\alpha/\alpha| < 3\times 10^{-6}$ at redshifts less than
1 \cite{Drietal98} and $|\Delta\alpha/\alpha| \mathrel{\mathpalette\fun <} 3\times
10^{-4}$ at redshifts of 3 \cite{CowSon95,IvaPotVar98}. In fact,
a detection of $\Delta\alpha/\alpha = -1.9\pm 0.5 \times
10^{-5}$ has been claimed on the basis of absorption lines at
redshifts greater than 1 \cite{Webetal99}, but there are some potential
problems with this result
\cite{IvaPotVar98}. Primordial nucleosynthesis can also provide
a less useful model-dependent limit \cite{KolPerWal86}.
A change in $\alpha$ would affect the recombination
rate of hydrogen and thus alter the redshift of
last scatter.
This effect on the CMB can potentially lead to
upper limits on
$|\Delta\alpha/\alpha|$ between 0.01 and 0.001
\cite{KapSchTur98,Han98a} out to redshifts $z\simeq1100$, much
larger than those probed by quasar absorption spectra.
\subsection{Topology of the Universe}
The fundamental cosmological assumptions of homogeneity and
isotropy require the Universe to be either the open, closed, or
flat Friedmann-Robertson-Walker model. However, if the
assumption of isotropy is incorrect, then the Universe may
have some nontrivial topology (see~\cite{LacLum95}
for a review).
The open and flat FRW models have infinite volume, but a
Universe with either zero or negative curvature can have finite
volume if the Universe has nontrivial topology. A number
of (somewhat imprecise) theoretical arguments suggest that a finite
Universe is easier to explain than an open Universe
\cite{ZelSta84} or could be used to explain the homogeneity of
the Universe \cite{Got80,EllSch86}.
If the volume of such a Universe is comparable to or less than that
observable today, then there may be signatures in the
CMB. Consider the simplest nontrivial topology (for a
flat Universe), that of a toroid. If the Universe is a
three-dimensional toroid, then two different directions on the
sky will point to the same point in space, and there should be
observable correlations between the CMB temperature at distant
locations on the sky. Such models have essentially been
ruled out by \COBE\
\cite{FanMo87,Sok93,Sta93,Fan93,SteScoSil93,JinFan94,CosSmo95,CosSmoSta96,ScaLevSil98,LevScaSil98}.
Interest in negative-curvature models with nontrivial topology
has reawakened recently because evidence seems to suggest
$\Omega_0\simeq0.3<1$, and thus possibly an open Universe.
If the Universe is negatively curved (hyperbolic), the spacetime
volume element increases rapidly with distance, so that even if
the volume of the Universe is close to the horizon volume,
many copies of the Universe may still fit inside the
horizon volume. Thus, none of the flat-Universe limits
on topology apply to hyperbolic
Universes \cite{CorSpeSta98b}. Furthermore, if the total density of
the Universe is $\Omega\simeq0.3$, the curvature scale is
small enough so that a huge number of topologies exist that
have proper volumes significantly smaller than the proper Hubble
volume \cite{CorSpeSta98c}.
Because the surface of last scatter is spherical, matched pairs of
temperature circles would
appear in a negatively-curved Universe
with nontrivial topology provided that the topology radius were
smaller than the current horizon
\cite{CorSpeSta96,Wee98,CorSpeSta98c}.
Levin et al \cite{Levetal98}
propose searching for specific correlations between a given
pixel and all others in a map. A null search for such
correlations in the \COBE\ maps ruled out a particular horn
topology \cite{Levetal97}. Souradeep et al \cite{SouPogBon98} claim that
the \COBE\ maps already rule out most hyperbolic Universes
through this technique, although details have not been
presented.
\subsection{Primordial Magnetic Fields}
Magnetic fields of strength $10^{-6}$ G are
ubiquitous in our Galaxy and in distant clusters of galaxies.
All mechanisms for the origin of these magnetic fields postulate
that they grew via some mechanism (e.g.\ dynamo or adiabatic
compression) from small
primordial magnetic fields. However, the origin of these
primordial seed fields remains a mystery. Many of the most
intriguing hypotheses the origin of these fields come from
new ideas in particle theory. Proposed generation mechanisms
include inflation
\cite{TurWid88,CarFie91,GarFieCar92,Rat92,Dol93,GasGioVen95a,GasGioVen95b},
the electroweak \cite{Vac91,EnqOle93} or QCD phase transitions
\cite{QuaLoeSpe89,CheOli94}, a ferromagnetic Yang-Mills vacuum
state \cite{EnqOle94}, charge asymmetry \cite{DolSil93}, and
dilaton evolution \cite{Gio97}.
Magnetic fields have several potentially
measurable effects: Faraday rotation \cite{KosLoe96} (A Mack,
A Kosowsky, manuscript in preparation) and
and associated depolarization \cite{HarHayZal96} of the
original CMB polarization; magnetosonic
waves that modify the acoustic oscillation frequencies
\cite{Adaetal96}; and Alfven waves, which can amplify vector
perturbations and induce additional correlations
\cite{DurKahYat98}, and for which diffusion
damping is decreased, thereby increasing CMB power at
small scales \cite{SubBar98}.
The Faraday rotation signals can be detected through
the CC, TC, and GC
power spectra they induce \cite{ScaFer97}
(although these power spectra are
frequency dependent).
A recent analysis of the \COBE\ maps has placed a limit on
a homogeneous primordial field strength corresponding
to $B_0< 3.4\times 10^{-9}(\Omega_0h_{50}^2)^{1/2}$ G
\cite{BarFerSil97} by searching for the temperature pattern of
a Bianchi type VII anisotropic spacetime
\cite{Nov68,BarJusSon85}.
\subsection{Large-Scale Parity Violation}
It is usually assumed that gravity is parity-invariant.
However, weak interactions are parity-violating
\cite{LeeYan56,Wu57}, and we surmise that the electroweak
interactions are united with gravity at the Planck scale by some
fundamental unified theory. Are there any
remnants of parity-violating new physics in the early Universe?
As discussed in Section \ref{sec:powerspectra}, if either of
the temperature-polarization cross-correlation moments $C_l^{\rm
TC}$ or $C_l^{\rm TG}$ is nonzero, it signals cosmological
parity breaking.
Lue et al and Lepora \cite{LueWanKam98,Lep98}
discuss how a parity-violating term, $\phi F_{\mu\nu}\tilde F^{\mu\nu}$
\cite{CarFie90,CarFie91,Car98}, that
couples a scalar field $\phi$ to the pseudoscalar ${\vec E}\cdot
{\vec B}$ of
electromagnetism could yield nonzero $C_l^{\rm TC}$ and
$C_l^{\rm GC}$. Lue et al \cite{LueWanKam98} also discuss a
parity-violating term in the Lagrangian for gravitation that
would yield an asymmetry between the density of right- and
left-handed gravitational waves produced during inflation; such
an asymmetry would also give rise to nonzero $C_l^{\rm TC}$ and
$C_l^{\rm GC}$. These parity-breaking effects would produce
frequency-independent $C_l^{\rm TC}$ and $C_l^{\rm GC}$,
unlike the frequency-dependent effect of Faraday rotation.
\subsection{Baryon Asymmetry}
There are very good theoretical and observational reasons to
believe that
our entire observable Universe is made of baryons and no
antibaryons. But suppose momentarily that the observable
Universe consisted of some domains with
antibaryons rather than baryons (see e.g.\ \cite{Ste81}).
If so, then particle-antiparticle
annihilations at the interfaces of the matter and antimatter
regions would release a significant amount of energy
in $\gamma$-rays, thus heating the region and causing a
$y$-distortion of the CMB spectrum of order $y\simeq 10^{-6}$
\cite{KinKolTur97,CohRujGla98}. These distortions would appear in
thin strips on the sky which could potentially be
identified. However, the point is moot because limits on
the diffuse extragalactic $\gamma$-ray background limit the size of
our matter domain to be essentially as large as the horizon
\cite{CohRujGla98}.
\subsection{Alternative Gravity Models}
We now know through a variety of experiments
that general relativity provides an accurate accounting of
observed gravitational phenomena. On the other hand, string theories
generically predict at least some small deviation from general
relativity, often in the form of scalar-tensor theories of
gravity \cite{Ber68,Nor70,Wag70,Bek77,BekMei78}. The simplest
of these is
Jordan-Fierz-Brans-Dicke (more commonly,
Brans-Dicke) theory
\cite{Jor49,Fie56,Jor59,BraDic61,Dic62,Dic68}. An inflation
theory (``extended inflation'') based on Brans-Dicke gravity
\cite{LaSte89a,LaSte89b} was ruled out by the isotropy of the
CMB \cite{Wei89,LaSteBer89}, although models based on more
complicated scalar-tensor theories (``hyperextended inflation'')
have also been considered
\cite{SteAcc90,BarMae90,GarQui90,HolKolWan90}.
Brans-Dicke theory includes a scalar field $\Phi$ and a new parameter
$\omega$. As $\omega \rightarrow \infty$, the theory recovers general
relativity (in some sense). Solar-system constraints from
Viking spacecraft data limit $\omega \geq 500$ (for a
review, see \cite{Wil93}) and recent Very-Long Baseline Interferometry
measurements of time
delays of millisecond pulsars may further raise this limit
\cite{Wil98}. In cosmological models based on Brans-Dicke
theories, general relativity is an attractor solution
\cite{DamNor93a,DamNor93b}, so gravity could have conceivably
differed from general relativity in the early Universe even if
it resembles general relativity today.
Because the expansion rate and growth of gravitational-potential
perturbations are different in alternative-gravity theories,
the precise predictions for CMB power spectra should be
different in these models.
The epoch of matter-radiation equality is altered in
Brans-Dicke theories, and this may produce an observable signal
in forthcoming precise CMB maps \cite{LidMazBar98}.
Cosmological perturbation
theory in scalar-tensor theories has been worked out
\cite{Nar69,PeeYu70,BapFabGon96,ChiSugYok98} and the CMB power
spectra calculated
(X Chen, M Kamionkowski, manuscript in preparation).
If the scalar-field time
derivative $\dot\Phi$ is fixed to be small enough to be
consistent with big-bang-nucleosynthesis constraints
\cite{KamTur90,DamGun91,CasGarQui92,DamPic98} and
$\omega>500$, then the differences between the
general-relativistic and Brans-Dicke predictions is small,
although conceivably detectable with the Planck Surveyor.
Of course, the scalar-field evolution may be significantly
different in more sophisticated scalar-tensor theories, but
predictions for these models have yet to be carried out.
\subsection{Cosmic Rays}
We close this tour of the CMB/particle intersection
with possibly the oldest and most venerable
connection between these two topics. Soon after the initial
discovery of the CMB, it was realized that cosmic rays
with energy $E\ga5\times 10^{19}$ eV can scatter from CMB
photons and produce pions. If a cosmic ray
is produced with an energy above $5\times 10^{19}$ eV, repeated
scatterings will reduce its energy to below this threshold
within a distance of about 50 Mpc
\cite{Gre66,ZatKuz66,Cro92,ElbSom95} (the
Greisen-Zatsepin-Kuzmin bound).
These constraint have become increasingly intriguing recently, as
several cosmic rays with energies $>10^{20}$
eV have been observed \cite{Lin63,Biretal94,Hayetal94,Biretal95}, and they do not appear to be coming from any
identifiable astrophysical sources (e.g.\ radio
galaxies or quasars \cite{Hil84}) as near as 50 Mpc
\cite{Hayetal94,Biretal95,ElbSom95,Bie97}. So where are these
cosmic rays coming from? Some possibilities are exotic
production mechanisms such as topological defects
\cite{Hiletal86,AhaBhaSch92,BhaHilSch92,Chietal93,
SigSchBha94} or supermassive unstable particles
(\cite{KuzRub98,BerKac98,BirSar98}; see \cite{SigBha98} for
a review). If a recently claimed alignment of the
highest-energy events with very distant radio quasars
\cite{FarBie98} is confirmed by larger numbers of events, then it
may be that these cosmic rays are exotic particles that interact
with baryons but not photons \cite{ChuFarKol98,AlbFarKol99},
e.g.\ supersymmetric $S_0$ baryons \cite{Far84,Far95,Far96}.
In the absence of any compelling traditional astrophysical
origin, it seems that the simultaneous existence of the CMB and
these cosmic rays may be pointing to some intriguing new
particle physics.
\section{Summary, Current Results, and Future Prospects}
The primary cosmological observables pursued by CMB experiments
are the frequency spectrum of the CMB, parameterized by
$\mu$ and $y$ distortions, and the angular temperature and
polarization power spectra, $C_l^{\rm TT}$, $C_l^{\rm GG}$,
$C_l^{\rm CC}$, $C_l^{\rm TG}$, $C_l^{\rm TC}$, and $C_l^{\rm
GC}$. There are additional observables, such as higher-order
correlation functions or cross-correlation of the CMB
temperature/polarization with other diffuse extragalactic
backgrounds. Rough
estimates of the $C_l^{\rm TT}$ at degree angular scales were
obtained \cite{WhiScoSil94} from the first generation of ground-based and
balloon-borne CMB experiments. Forthcoming
experiments will require far more sophisticated techniques for
disentangling the CMB from foregrounds, and for recovering the power
spectra from noisy data and from maps that cover only a fraction
of the sky. A large literature
is now devoted to these important issues,
which we cannot review here.
Progress in CMB experiments is so rapid at the time
of writing that any current data we might review would
almost certainly become obsolete by the time of publication.
We therefore refrain from showing any experimental results
in detail and instead describe the current observations qualitatively.
First, there is the isotropy of the CMB, which has long been
explained only by
inflation. Among the numerous pre-\COBE\ models for the origin
of large-scale structure, those based on a nearly
scale-free spectrum of primordial adiabatic perturbations seem
to account most easily for the amplitudes of both the
large-angle CMB anisotropy measured by \COBE\ and the amplitude
of clustering in galaxy surveys. The galaxy distribution seems
to be consistent with primordial Gaussianity.
Moreover, data from a large number of CMB experiments that probe
the angular power spectrum at degree angular scales have now
found (fairly convincingly) that there is significantly more
power at degree angular separations ($l\sim200$) than at \COBE\
scales, as one would expect if the acoustic peaks do
exist, but in apparent conflict with most theorists'
expectations for the degree-scale anisotropy in
topological-defect models.
The
existence of this small-scale anisotropy further suggests no
more than a small level of reionization (i.e.\ $\tau\ll 1$). At
the time of writing, the measurements are not precise
enough to discern either the first or any higher peaks in the
temperature power spectrum (some recent data are shown in
Reference \cite{Kam98} and are usually updated at Reference \cite{Teg99}).
Some experiments have claimed to see the outline of a first
acoustic peak at $l\sim200$ \cite{Petetal99} (which would
indicate a flat Universe). Moreover, some maximum-likelihood
analyses of combined results of all experiments claim that the
data indicate a flat Universe \cite{LinBar98,Hanetal98}.
However, these results are not yet robust.
Thus, although inflation is by no means yet in the clear,
observations do seem to be pointing increasingly toward
inflation. MAP and the Planck Surveyor will soon make far more
precise tests of inflation (see \cite{Kam98} for simulated
data from MAP and Planck). First of all, the predictions of
primordial adiabatic perturbations will be tested with
unprecedented precision by the peak structure in the CMB
temperature power spectrum. If the peaks do appear, then MAP
and the Planck Surveyor should be able to measure the total
density $\Omega$ to a few percent or better \cite{Junetal96a} by
determining the location of the first acoustic peak
\cite{KamSpeSug94}. Moreover, by fitting MAP and Planck
satellite data to theoretical curves, such as those shown in
Figure \ref{fig:models}, precise information on the values of
other classical cosmological parameters can also be obtained
\cite{Junetal96b,BonEfsTeg97,ZalSelSpe97,DodKinKol97,BonEfs98}.
If nonrelativistic matter outweighs baryons, then it should
become evident with MAP and Planck. The existence of a
cosmological constant will further be tested, and some of the
tests of gravity, decaying particles, etc, that we
have reviewed will become possible.
If MAP and Planck confirm that the Universe is flat and that
structure grew from primordial adiabatic perturbations, then the next
step will be to search for the gravitational-wave background
predicted by inflation. Such a gravitational-wave background
could be isolated uniquely with the curl component of the
polarization. If the inflaton-potential height is $V^{1/4}\ll
10^{15}$ GeV, then the gravitational-wave background will be
unobservably small. However, if inflation had something to do
with grand unification (i.e.\ $V^{1/4} \sim 10^{15-16}$~GeV, as
many theorists surmise), then the curl component of the
polarization is conceivably detectable with the Planck Surveyor
or with a realistic next-generation dedicated polarization
satellite experiment. If detected, the curl component would
provide a ``smoking-gun'' signature of inflation and indicate
unambiguously that inflation occurred at $T\sim10^{15-16}$ GeV.
Although an observable signature is by no means guaranteed, even
if inflation did occur, the prospects for peering directly back to
$10^{-40}$ sec after the big bang are so
tantalizing that a vigorous pursuit is certainly warranted.
\bigskip
\leftline{\textsc{Acknowledgments}}
We thank R Caldwell, A Liddle, and L Wang for very
useful comments.
MK was supported by a DOE Outstanding Junior Investigator
Award, DE-FG02-92ER40699, NASA Astrophysics Theory Program grant
NAG5-3091, and the Alfred P. Sloan Foundation. AK was supported
by NASA Astrophysics Theory Program grant NAG5-7015 and
acknowledges the kind hospitality of the Institute for Advanced
Study.
{\twocolumn
|
\section{Introduction}
According to the theory of strong interactions, Quantum
Chromodynamics (QCD)~\cite{qcd}, the strong coupling constant \as\
is the same for all quark flavours. Therefore a precise
measurement of \as\ for the individual quark flavours is an
important test of this theory. The coupling constant \as\
for charm and bottom
quarks can be compared to \as\ for light (uds) quarks
by measuring the ratios \rascasu\ and \rasbasu\ using
hadronic events of the type $\epem \rightarrow \qqbar g$.
The coupling constants are measured
using an event sample originating from a pair of light
quarks (\uubar , \ssbar\ or \ddbar),
c quarks (\ccbar) or b quarks (\bbbar), respectively.
If either of the above ratios
deviates significantly from unity then this may indicate
physics beyond the Standard Model.
In many QCD studies at LEP corrections due to quark mass effects
can be safely ignored
because they typically appear as powers of the ratio
of the quark mass to the total energy. However, in those studies
where \as\ is determined in event samples enriched in heavy
quarks, mass effects become non-negligible.
Gluon emission by bottom quarks, and
to a lesser extent charm quarks, will be suppressed largely due
to the reduced phase-space available.
Observables sensitive to the three-jet rate
measured in heavy quark events will be modified with respect
to the same quantities measured in light quark events.
Tests of the flavour independence of \as\ have previously been
conducted at both LEP and SLC
~\cite{flvtstopal1,flvtstopal2,flvtstl3,flvtstdelphi,flvtstaleph,flvtstsld}
using ratios of \as\ for one flavour over \as\ of either the
complementary
\footnote{For example, \as\ for c quarks was
compared to \as\ for a mixture
of u, d, s and b quarks.}
or inclusive
quark mixture. In all these tests flavour independence of \as\
was confirmed. These tests, however, used either massless QCD
calculations or a leading order calculation with massive quarks~\cite{heavyQ0}.
The leading order calculation includes
the process $\epem \rightarrow \qqbar gg$,
with mass effects, but virtual
corrections to the process $\epem \rightarrow \qqbar g$
are not included.
Complete next-to-leading-order (NLO) calculations
of the heavy-flavour production cross section in \epem\ collisions,
including quark mass effects,
have been published recently~\cite{heavyQ1, heavyQ2, heavyQ3}.
A comparison between these calculations
has been made and they were found to be in agreement~\cite{heavyQ3}.
In recent publications DELPHI~\cite{delphi_massive} and
SLD~\cite{sld_massive} report
on measurements of \rasbasu\ where NLO
massive calculations~\cite{heavyQ1,heavyQ2} were used to account for
mass effects in b quark events. These results are consistent
with the flavour independence of \as .
In this study we use the
results of P.~Nason and C.~Oleari~\cite{heavyQ3}
along with theoretical
predictions assuming massless quarks in fits to
global event shape distributions
in order to determine \asu\ , \rascasu\ and \rasbasu.
The results presented here are intended to update and
supersede the corresponding OPAL results in~\cite{flvtstopal1,flvtstopal2}
insofar as we now have greatly increased the charm event statistics and
have used improved theoretical predictions.
This paper is organised as follows. In Sect.~\ref{sec:dete}
the parts of the OPAL detector most important to this analysis
are described. In Sect.~\ref{sec:evt} the hadronic event sample
and the Monte Carlo event sample are introduced.
In Sect.~\ref{sec:buds} and \ref{sec:dstar} the
flavour tagging methods are
described. In Sect.~\ref{sec:shp} the event shape
observables used in this study are
introduced and the procedure for correcting the
event shape distributions is explained.
Next, in Sect.~\ref{sec:oas2}, the procedure for fitting the
NLO QCD prediction to the corrected distributions is explained and
in Sect.~\ref{sec:sys} the systematic uncertainties that
have been taken into account are
discussed. In Sect.~\ref{sec:res},
the results of the test of flavour independence of \as\ are
presented.
Finally, in Sect.~\ref{sec:con}, conclusions are drawn.
\section{The OPAL detector}
\label{sec:dete}
The OPAL detector operates at the LEP \epem\ collider at CERN. A
detailed description can be found in
Refs.~\cite{opaldete,si}. The analysis presented here
relies mainly on the reconstruction of charged particle trajectories
and momenta in the
central tracking chambers, on energy deposits (``clusters'') in
the electromagnetic calorimeters and on information from
the silicon micro-vertex detector.
All tracking systems are located inside a solenoidal magnet which
provides a uniform magnetic field of 0.435~T along the beam
axis\footnote{In the OPAL coordinate system the $x$ axis points
towards the centre of the LEP ring, the $y$ axis points upwards and
the $z$ axis points in the direction of the electron beam. The
polar angle $\theta$ and the azimuthal angle $\phi$ are defined
with respect to the $z$- and $x$-axes, respectively, while $r$
is the distance from the
$z$-axis.}.
The magnet is surrounded by a lead glass electromagnetic
calorimeter and a hadron calorimeter of the sampling type. Outside
the hadron calorimeter, the detector is surrounded by a system of
muon chambers. There are similar layers of detectors in the barrel
($|\rm{cos}\theta | < 0.82$) and endcap ($|\rm{cos}\theta | > 0.81$)
regions.
The central tracking detector consists of a silicon micro-vertex
detector~\cite{si} and three drift chamber devices: the vertex
detector,
a large jet chamber, and surrounding $z$-chambers. The silicon
micro-vertex detector, close to the beam pipe, consists of two
layers of silicon strips with a single-hit resolution of about
7\,$\mu$m in the $r\phi$ plane.
The vertex chamber is a cylindrical drift chamber covering
a range of $|\cos\theta|<0.95$. Its single hit resolution is
50\,$\mu$m in the $r\phi$ plane and 700\,$\mu$m in the $z$ direction.
The jet chamber is a cylindrical drift chamber with an inner radius
of 25\,cm, an outer radius
of 185\,cm, and a length of about 4\,m.
Its spatial resolution is about 135\,$\mu$m in the $r\phi$ plane
from drift time information and about 6\,cm in the $z$ direction
from charge division. The $z$-chambers
provide a more accurate $z$
measurement with a resolution of about 300\,$\mu$m. In combination,
the three drift chambers yield a momentum resolution of
$\sigma_{p_t}/p_t \approx \sqrt{0.02^2+(0.0015\cdot p_t)^2}$
for $|\cos(\theta)| < 0.7$, where $p_t$ is the transverse momentum
in GeV/$c$.
Electromagnetic energy is measured by lead glass calorimeters
surrounding the solenoid magnet coil. They consist of a barrel and
two endcap arrays
with a total of 11704 lead glass blocks covering a range of
$|\cos\theta|<0.98$.
\section{Event sample and Monte Carlo simulation}
\label{sec:evt}
This analysis is based on a sample of 4.4 million hadronic
decays of the \znull\ recorded with the OPAL
detector between 1990 and 1995.
Hadronic \znull\ decays were selected by placing requirements on
the number of reconstructed tracks and the energy deposited
in the calorimeter. A detailed description of the criteria
is given in~\cite{bib-OPALmh}.
The parts of the detector essential for the present analysis
(central detector and electromagnetic calorimeter) were required to be
fully operational.
The track selection criteria were the same as presented in a
previous OPAL study~\cite{opalresummed}.
The number of accepted
tracks was required to be at least five to reduce $\tau^+\tau^-$
background. Clusters of electromagnetic
energy were used if their observed energy was greater than 0.25\,GeV,
and known noisy channels in the detector were ignored.
The event thrust axis~\cite{opalresummed} was determined using
all accepted tracks and clusters, and its direction was
required to fulfil the condition $|\cos\theta_\mathrm{Th}|<0.9$ in
order that the event be well contained. Using these
selection criteria, Monte Carlo studies indicate that, within the
chosen range of $\cos\theta_\mathrm{Th}$,
99.86$\pm$0.07\% of hadronic \znull\ decays are accepted, with a
contamination
of about 0.14\% from $\tau^+\tau^-$ events, and around 0.07\% from
two-photon interactions~\cite{opalresummed}.
To correct the measured event shape distribitions (see
Sect.~\ref{sec:cor}), 4 million
hadronic decays of the \znull\ have been simulated using the
JETSET 7.4 Monte Carlo model~\cite{bib-JETSET}
with parameters tuned to represent OPAL data well
\cite{bib-OPALtune}. For all simulated events heavy quark fragmentation
has been implemented using the model of
Peterson et al.~\cite{bib-PETERSON}.
All events have been passed through a detailed
simulation of the OPAL detector~\cite{bib-OPALGOPAL}
before being analysed using the same programs as for data.
\section{Selection of uds and b quark events}
\label{sec:buds}
Two event samples, one enriched in uds quark events and one
enriched in b quark events, were selected by first determining the number
of tracks in each event ($N_{sig}$) with a large impact parameter
significance, $b/\sigma_b > 2.5$. Here $b$ is the distance of closest approach in
the $xy$ plane of the track
to the \epem\ interaction point (IP) and $\sigma_b$ its error.
The sign of $b$ was determined with respect to the crossing
point between the track and the jet axis.
$b$ is positive if the track crosses the jet axis
downstream of the IP and negative otherwise.
An event was classified as having a high probability
to have come from light quarks
if $N_{sig}$ was zero and from b quarks if
$N_{sig} \geq 5$.
In the text below these will be referred to as the ``uds-tag'' and
the ``b-tag'' event samples.
For the uds and b quark event selection all data recorded
during 1994 were used. This represents 1.4 million events.
The analysis was restricted to only the 1994 data because of the
uniform configuration of the silicon micro-vertex detector
during this time period.
In addition to the track selection criteria outlined above,
the track was
required to contain at least one silicon hit and
an algorithm was applied to reject tracks which were
consistent with arising from photon conversions~\cite{bib-idgcon}.
In order for the event to be contained within the acceptance
of the silicon micro-vertex detector
the event's thrust axis was
further restricted to lie within the range
$|\cos\theta_\mathrm{Th}| < 0.7$.
The distribution of $N_{sig}$ is shown in Fig.~\ref{fig:nsig}
compared with the result of Monte Carlo simulation.
The simulation is decomposed into the contributions from uds, c
and b quark events. The general agreement between the sum of the
three contributions from Monte Carlo and the data is good.
This tagging procedure resulted in 325\,111 events selected for
the uds-tag event sample and 71\,521 for the b-tag event sample.
The flavour compositions of the uds-tag and the b-tag event samples
as determined from the Monte Carlo, along with their combined
statistical and systematic errors,
are presented in Table~\ref{tab:flavor}.
The efficiency for tagging uds events was about
35\% and for tagging b events 23\%.
\subsection{Uncertainties in Flavour Composition}
The systematic errors on the flavour compositions result from
uncertainties in the detector modelling and imprecise knowledge
of physics processes. For each source of uncertainty some
aspect of the Monte Carlo was varied and the flavour
composition of the uds-tag and b-tag event samples were
recalculated. The difference between the flavour fractions
calculated for each variation and the central value was taken
as a systematic error on the determined flavour fractions.
\subsubsection{Detector Modelling Uncertainties}
Since the determination of the flavour fractions of the uds-tag
and the b-tag event sample depend on Monte Carlo this requires
an accurate simulation of the detector resolution for charged
tracks measured with the silicon micro-vertex detector.
The simulation has been tuned to reproduce the
tracking resolutions seen in data by studying the impact parameter
distributions of tracks, as functions of track momentum,
polar angle and the different sub-detectors contributing
hits. This tuning procedure was affected by uncertainties in
the radial alignment within the silicon micro-vertex detector, the
efficiency for associating silicon hits to tracks and the modelling
of known inefficient regions, and the overall track reconstruction efficiency.
These uncertainties were evaluated according to the procedure given in
\cite{opalrb} and their effect on the flavour fractions determined
from Monte Carlo was calculated.
In addition, the agreement between data and Monte Carlo in
Fig.~\ref{fig:nsig} was improved by degrading the resolution of
impact parameters in the Monte Carlo simulation
by 5\%. This was done by applying a single multiplicative
factor $\beta$ to the difference between the reconstructed
and true impact parameters.
Fig.~\ref{fig:nsig} is shown with this smearing applied.
To evaluate the sensitivity of the
determined flavour fractions to the tracking resolution the
Monte Carlo smearing was removed.
\subsubsection{Physics Modelling Uncertainties}
The tagging efficiencies of the uds and b tags for the various quark flavours
are also sensitive to various physics input parameters in the
Monte Carlo simulation. The rate of gluon splitting to \ccbar\ was varied
in the range $(2.38\pm 0.48)\times 10^{-2}$, based on the OPAL measurement
\cite{opal_glue_split}, and the rate of gluon splitting to \bbbar\ was varied in the
range $(3.1\pm 1.1)\times 10^{-3}$, based on theoretical expectation
\cite{glue_split_calc}. The error due to these two sources was negligible.
The production fractions of the different weakly decaying
b hadrons were varied according to the experimental uncertainties
\cite{PDG}. The production fractions of the weakly decaying c hadrons,
the fragmentation of b and c quarks, and the charged decay multiplicities
and lifetimes of b and c hadrons were varied according to the
prescription given in \cite{opalrb}. The largest effect on the tag flavour
fractions comes from the b hadron charged decay multiplicity. All these
uncertainties, along with the uncertainties
due to the detector modelling, were added in quadrature and are given as
the errors on the flavour fraction shown in Table~\ref{tab:flavor}.
\section{Selection of c quark events}
\label{sec:dstar}
Events having a high probability to
have originated from \ccbar\ were identified by the presence
of a highly energetic D$^{*+}$ meson\footnote{Throughout this paper
charged conjugate modes are always implicitly included.}.
These events will be referred to in the text below as the ``c-tag''
event sample.
For the D$^{*+}$ reconstruction and the determination
of the flavour composition of the c-tag event sample
the methods of a previous
OPAL study have been used~\cite{dstarsel}.
In brief, five D$^{*+}$ decay modes were searched for:
\begin{center}
\begin{tabbing}
\hspace{5cm} \= \hspace{5cm} \= \kill
\> ${\rm D^{*+}} \rightarrow {\rm D^0}\pi^+$ \\
\> $\phantom{D^{*+} \rightarrow }\hspace{4pt}
\downto {\rm K^-}\pi^+$ \>``3 prong''\ ,\\
\> $\phantom{D^{*+} \rightarrow }\hspace{4pt}
\downto {\rm K^-}\pi^+\pi^0$ \> ``satellite''\ ,\\
\> $\phantom{D^{*+} \rightarrow }\hspace{4pt}
\downto {\rm K^-}\pi^+\pi^-\pi^+$ \>``5 prong''\ ,\\
\> $\phantom{D^{*+} \rightarrow }\hspace{4pt}
\downto {\rm K^-}{\rm e}^+ \nu_{{\rm e}}$ \>``electron''\ ,\\
\> $\phantom{D^{*+} \rightarrow }\hspace{4pt}
\downto {\rm K^-}\mu^+\nu_{\mu}$ \>``muon''\ .\
\end{tabbing}
\label{eq-decaychannels}
\end{center}
\noindent
No attempt was made
to reconstruct the $\pi^0$ in the satellite channel, nor the
neutrino direction or energy in the electron and muon channels.
The last two channels are referred to as ``semileptonic'' channels
in the following text. A number of tracks
appropriate for the selected channel were
combined to form a D$^0$ candidate and their invariant mass
was calculated.
Candidates were selected if the reconstructed mass lay
within the expected range for that channel.
After adding a further track
as a possible pion from the D$^{*+}$ decay, the
combined mass was calculated and the candidate was selected
if the mass difference $\Delta M = M_{D^{*+}}-M_{D^0}$
was within certain limits.
Some of the D$^{*+}$ selection criteria are given in
Table~\ref{tab:dstarcuts}.
To reduce the background in the c-tag event sample we required
$x_{{\rm D^{*+}}} > 0.4$, where
$x_{{\rm D^{*+}}}=E_{{\rm D^{*+}}}^{\mathrm calc}/E_{\rm beam}$
is the scaled
energy\footnote{In this paper any reference to the scaled energy
$x$ of a D$^{*+}$ candidate is to be understood as
being the calculated energy of the D$^{*+}$,
$E_{{\rm D^{*+}}}^{\mathrm calc}$, obtained from the
reconstructed
tracks, without correcting for missing or wrongly
associated tracks, divided by the beam energy
$E_{\mathrm beam}$.}
of the D$^{*+}$ meason. For the 5-prong event selection we required
$x_{{\rm D^{*+}}} > 0.5$. This cut is effective in rejecting
D$^{*+}$ mesons which originate from cascade decays
of B hadrons and from events where a gluon splits into
a pair of charm quarks.
The distribution of $\Delta M$ for all five decay channels
is presented in Fig.~\ref{fig:delm}. The points with error bars
are the signal candidates and the solid histograms are
background estimator distributions constructed from data.
The background estimator was constructed by choosing the
candidate for the pion in the
$ {\rm D^{*+}} \rightarrow {\rm D^0}\pi^+$
decay from the opposite hemisphere relative to the rest
of the decay products, reflecting it through the orgin,
and then using it in the calculation of the invariant mass.
A significant fraction of the sample of D$^{*+}$ mesons were only partially
reconstructed. These mesons produce an enhancement in the
$\Delta M$ spectrum very similar to the true signal. Only a few
of the events are present in the 3-prong sample. They are more important in
the 5-prong tagged events where a clear tail is visible in the $\Delta M$
distribution for values above 0.145 GeV (see Fig.~\ref{fig:delm}d).
Since such events originate from D$^{*+}$ decays, they can still be used in
the analysis.
In all, 27\,005 D$^{*+}$ candidate events were selected and the
background was estimated to be $11\,366 \pm 107$ events, where
the error on the number of background events is
statistical only. All of the LEP-1 data recorded by OPAL,
4.4 million events, was used for the D$^{*+}$ event selection.
The candidate events are composed of three
components: genuine D$^{*+}$ mesons from b events,
genuine D$^{*+}$ mesons from c events and combinatorial
background which is a mixture of uds, c and b quark events.
Genuine D$^{*+}$'s from uds quark events can only occur in
events where a gluon splits into two heavy quarks.
The possibility of D$^{*+}$'s from this source was neglected because
these events are highly suppressed due to the
high $x_{\rm D^{*+}}$ cut. Using Monte Carlo simulation
the total contribution of this source to the number
of D$^{*+}$ candidates was found to be $(0.2 \pm 0.1)$\%, where the
error quoted is due to Monte Carlo statistics.
In Ref.~\cite{dstarsel} the fraction of genuine D$^{*+}$
mesons originating from c events was determined by OPAL to be
$f_{\rm c}^{{\rm D^{*+}}} = 0.774 \pm 0.023$, where the error is the
combined statistical and systematic error. The fraction of
genuine D$^{*+}$ mesons from b events is given by
$f_{\rm b}^{{\rm D^{*+}}} = 1 - f_{\rm c}^{{\rm D^{*+}}}$.
The fractions of uds, c and
b quark events in the combinatorial background, as determined
from Monte Carlo, were
$0.584 \pm 0.009$, $0.238 \pm 0.009$ and
$0.178 \pm 0.009$, respectively. The errors are a combination
of statistical errors due to finite Monte Carlo statistics
and a systematic error accounting for the
overall quality of the background estimation
procedure. Based on a Monte Carlo study~\cite{dstarsel}
the background estimate was found to be accurate to within 1\%,
and an additional 1\% error was therefore
assigned to the number of background events.
Since that study was flavour blind, a conservative approach was
taken here where the entire error was assigned in turn to each
flavour component in the combinatorial background.
The overall flavour composition of the c-tag event sample is
presented in Table~\ref{tab:flavor} along with the combined
statistical and systematic errors. The efficiency for tagging
c events was about 2.0\%.
\section{Event shape observables}
\label{sec:shp}
For each flavour tagged event sample described
above (uds-tag, c-tag and b-tag)
the distributions of the event shape variables
1-Thrust ($1-T$), Heavy Jet Mass scaled by the
centre-of-mass energy ($M_H/\sqrt{s}$),
Wide Jet Broadening ($B_W$),
the $y_{cut}$ at which an event changes from
being a 2-jet event to being a 3-jet event
($y_{23}$) determined using the Durham jet
finder and the C-parameter ($C$) were determined.
The definitions for
these observables are given in~\cite{opalresummed, as_global}
and the references therein.
These quantities were measured using all tracks and
electromagnetic clusters which satisfied the
selection criteria described in Sect.~\ref{sec:evt}.
In order to extract the values of \rascasu\ and
\rasbasu , the measured event shape distributions were
fitted using a QCD analytic calculation~\cite{heavyQ3}.
Since the QCD calculation is only valid for event shape distributions
determined from final-state partons, hadronization effects
caused by the transformation of final-state partons into hadrons, which are
experimentally accessible,
must be taken into account. This was done
by applying correction factors to the analytic predictions.
The measured event shape distributions must be corrrected for
experimental effects which distort them.
These effects include finite detector resolution,
initial-state photon radiation and biases introduced by
the flavour tagging methods.
In the text below describing the correction procedure
the term {\it detector level} is used to refer
to distributions determined using the measured tracks and
electromagnetic clusters and {\it hadron level} to refer to
these distributions corrected for detector resolution,
initial state radiation and biases introduced by the flavour
tagging methods. The comparisons between the measured
distributions and the QCD predictions were performed
at the hadron level.
The event shape variable Total Jet Broadening
($B_T$)~\cite{opalresummed}
was initially considered to be included
in this analysis. It was found, however, that for the distribution of $B_T$
measured with b quark events, the size of the hadronization
corrections were greater that 20\%
over the entire distribution. Therefore
this variable was dropped from further consideration.
\subsection{Correction procedure}
\label{sec:cor}
In the first step the selected events
were corrected for distortions caused by finite
detector resolution using an unfolding matrix.
The unfolding matrix was constructed using
a Monte Carlo data sample including
initial-state radiation, full detector
simulation and subjected
to the same event selection criteria that were applied
to the data. The Monte Carlo events which pass the
selection criteria were used to calculate a
correction matrix $M^{q-tag}$.
The element $M^{q-tag}(y_i,y_j)$ gives the
probability that an event shape observable
$y$ measured at the detector level and located in
bin $i$ of its corresponding distribution,
has migrated from bin $j$ on the hadron level.
A matrix was computed
for each event shape observable and flavour-tagged
event sample,
``q-tag'', where q-tag was either uds-tag, c-tag or b-tag.
In the next step bin-by-bin corrections are used
to correct the data for biases introduced
by the flavour tagging methods, event acceptance and
the effects of initial-state photon radiation.
Defining $G^q(y_i)$ to be the number of events of
flavour $q$ in
the untagged Monte Carlo sample and $H^{q-tag}(y_i)$
the number of events with the tag applied,
the correction factor $K^{q-tag}(y_i)$ for the $i$-th bin is
given by
\begin{eqnarray}
K^{q-tag}(y_i) =
\frac{f^{q-tag}_{\rm uds} G^{\rm{uds}}(y_i) +
f^{q-tag}_{\rm c} G^{\rm{ c}}(y_i) +
f^{q-tag}_{\rm b} G^{\rm{ b}}(y_i) }
{H^{q-tag}(y_i)}
\end{eqnarray}
\noindent
The factors $f_q^{q-tag}$, taken from
Table~\ref{tab:flavor},
are the fractions of events of flavour $q$ in each flavour-tagged
event sample. The terms $G^q(y_i)$ are normalized to the total
number of events of flavour $q$.
Fig.~\ref{fig:bwcor} shows, as an example, the size of the bin-by-bin corrections
$K^{q-tag}$
for the observable $B_W$ measured using each flavour-tagged
event sample. The size of the corrections for
$B_W$ were typical of the other event shape observables studied.
One can see for the distributions
measured with the uds-tag and b-tag event samples that, within the
chosen fit range, the size of the corrections are about
10\%. (An explanation of how the fit range was chosen is presented in
the next section.) In contrast, the size of the corrections for the
c-tag event sample were an order of magnitude larger. This is due
to a kinematic
bias introduced by requiring the events in the
c-tag sample to contain a high-$x$ $D^{*+}$ meson.
The number of events in bin $i$ corrected to the hadron level is
given by
\begin{eqnarray}
N^{q-tag,cor}(y_i) = K^{q-tag}(y_i) \sum_{j} M^{q-tag}(y_i,y_j)N^{q-tag}(y_j)~~,
\end{eqnarray}
\noindent
where $N^{q-tag}(y_j)$ is the uncorrected number of events
in the $j$-th bin of the event shape distributions in question.
This correction procedure does not depend on
the values of \asu , \rascasu\ or \rasbasu\ in the
Monte Carlo samples.
The resulting distributions normalized to the total hadronic
cross section are given by
\begin{eqnarray}
\left(\frac{1}{\sigma_{tot}}\frac{\rm{d}\sigma}{\rm{d}y}\right)^{q-tag,cor}
= \frac{1}{\Delta y_i \cdot \sum_{j} N^{q-tag,cor}(y_j)} N^{q-tag,cor}(y_i)~~,
\end{eqnarray}
\noindent
where $\Delta y_i$ is the width of the $i$-th bin and $y$ corresponds
to one of the five event shapes studied.
\section{Fit procedure}
\label{sec:oas2}
The value of \asu\ and the ratios \rascasu\ and
\rasbasu\ were determined by simultaneously fitting theoretical
predictions for a particular event shape observable to three
hadron-level
event shape distributions: one determined with the uds-tag event sample,
one with the c-tag event sample and one with the b-tag event sample.
The theoretical predictions were given by a linear combination of the
theoretical predictions for each of the three tagged flavours:
\begin{eqnarray}
\label{eq:theo_shp}
\left(\frac{1}{\sigma_{tot}}\frac{\rm{d}\sigma}{\rm{d}y}\right)^{q-tag,th} &=&
\label{eqn:lincomb}
f_{\rm{uds}}^{q-tag}R(y)^{\rm{uds}}\left(\frac{1}{\sigma_{tot}}\frac{\rm{d}\sigma}{\rm{d}y}\right)^{{\rm uds},th} \nonumber
+ f_{\rm{ c}}^{q-tag}R(y)^{\rm{c }}\left(\frac{1}{\sigma_{tot}}\frac{\rm{d}\sigma}{\rm{d}y}\right)^{{\rm c },th} \\
& & \mbox{} + f_{\rm{ b}}^{q-tag}R(y)^{\rm{b }}\left(\frac{1}{\sigma_{tot}}\frac{\rm{d}\sigma}{\rm{d}y}\right)^{{\rm b },th}~~.
\end{eqnarray}
\noindent
The coefficients $f_q^{q-tag}$ are the flavour fractions
given in Table~\ref{tab:flavor} and
($1/\sigma_{tot}\cdot {\rm d}\sigma/{\rm d}y)^{\rm{uds},th}$,
($1/\sigma_{tot}\cdot {\rm d}\sigma/{\rm d}y)^{\rm{c},th}$ and
($1/\sigma_{tot}\cdot {\rm d}\sigma/{\rm d}y)^{\rm{b},th}$
are the theoretical predictions for an event shape observable
$y$ measured with a sample of uds, c and b quark events,
respectively.
The factors $R(y)^q$ correct
the theoretical prediction for
a particular tagged flavour $q$ for hadronization
effects so that the theoretical predictions, valid for
final state partons, can be compared
directly with the measured distributions
corrected to the hadron level.
The hadronization correction factors $R(y)^q$ were computed
from JETSET 7.4 using the
parton shower option
by taking the ratio of an event shape distribution at the
hadron level to the same distribution at the parton
level. Here the parton level is defined by the cut-off $Q_0$ of
the QCD shower in JETSET which is set to 1.9\,GeV~\cite{bib-OPALtune}.
After the termination of the parton shower the partons
are transformed into hadrons using string
hadronization~\cite{string-had}. For this process
we choose a hybrid scheme for the longitudinal fragmentation
function where light quarks are treated with the symmetric
Lund fragmentation function and charm and bottom quarks according to the
model of Peterson et al.~\cite{bib-PETERSON}.
The c and b quark masses in JETSET were set to their
default values of 1.35\,GeV and 5.0\,GeV, respectively.
In Equation~(\ref{eqn:lincomb}),
the differential cross section of a generic observable $y$,
for massless quarks,
normalized to the total hadronic cross section is
given by~\cite{oas2}
\begin{eqnarray}
\left(\frac{1}{\sigma_{tot}}\frac{{\rm d}\sigma}{{\rm d}y}\right)^{{\rm uds},th} & = &\nonumber
\label{eqn:udsth}
\frac{{\rm d}A^{\rm uds}}{{\rm d}y}\left(\frac{\alpha_s^{uds}(\mu)}{2\pi}\right) \\
& & \mbox{} + \left(\left(2\pi \beta_0 \mathrm{log}(x_\mu^2) - 2 \right)
\frac{{\rm d}A^{\rm uds}}{{\rm d}y} + \frac{{\rm d}B^{\rm uds}}{{\rm d}y}\right)
\left(\frac{\alpha_s^{uds}(\mu)}{2\pi}\right)^2~~.
\end{eqnarray}
\noindent
The coefficients
${\rm d}A^{\rm uds}/{\rm d}y$ and ${\rm d}B^{\rm uds}/{\rm d}y$ are the
\oa\ and \oaa\ QCD coefficients,
respectively, $\sigma_{tot}$ is the one-loop cross section
for the process $\epem \rightarrow \mathrm{hadrons}$ and
$\beta_0$ is the coefficient of the QCD beta function
for one-loop~\cite{PDG}. The renormalization scale
$\mu$ can be related to the \epem\ centre-of-mass
energy by
\begin{eqnarray}
\mu = x_\mu \cdot E_{cm}~~,
\end{eqnarray}
\noindent
where $x_\mu$ is the renormalization scale factor.
The coefficients ${\rm d}A^{\rm uds}/{\rm d}y$ and ${\rm d}B^{\rm uds}/{\rm d}y$
were obtained for each event shape observable
by integrating the \oaa\ matrix elements in~\cite{oas2}.
In Equation~(\ref{eqn:udsth}), in addition to the approximation
of massless quarks,
the simplifying assumption was made that \as\ is the same for
up, down and strange quarks.
The terms
($1/\sigma_{tot}\cdot {\rm d}\sigma/{\rm d}y)^{\rm{c},th}$ and
($1/\sigma_{tot}\cdot {\rm d}\sigma/{\rm d}y)^{\rm{b},th}$
in Equation~(\ref{eqn:lincomb}) were both given by an
\oaa\ expression for massive quarks.
This expression has only recently been made available~\cite{heavyQ3}.
The result of this calculation was implemented in a FORTRAN program
named ZBB4~\cite{zbb4} analogous to the program
EVENT~\cite{event} which was used to integrate the \oaa\
matrix elements of the massless calculation.
ZBB4 was
run separately for c quark and b quark events in order
to calculate the \oa\ and \oaa\ coefficients for
the massive calculation. This calculation was performed
in the pole mass scheme and
the c and b quark pole masses
were set to 1.35\,GeV and 5.0\,GeV, respectively.
The differential cross section of a generic observable $y$,
for massive quarks, normalized to the total hadronic cross section is
given by~\cite{zbb4}
\begin{eqnarray}
\left(\frac{1}{\sigma_{tot}}\frac{{\rm d}\sigma}{{\rm d}y}\right)^{Q,th} & = & \nonumber
\label{eqn:Qth}
\frac{{\rm d}A^{Q}}{{\rm d}y}\left(\frac{\alpha_s^{Q}(\mu)}{2\pi}\right) \\
& & \mbox{} + \left(\left(2\pi \beta_0 \mathrm{log}(x_\mu^2) -
\frac{2}{3}{\mathrm{log}}\left(\frac{\mu}{m_{Q}}\right) - 2 \right)
\frac{{\rm d}A^{Q}}{{\rm d}y} + \frac{{\rm d}B^{Q}}{{\rm d}y}\right)
\left(\frac{\alpha_s^{Q}(\mu)}{2\pi}\right)^2
\end{eqnarray}
\noindent
where $Q$ is either c or b and $m_Q$ is the energy scale corresponding
to the heavy quark pole mass.
When performing the fit using Equation~(\ref{eqn:lincomb}) we make
the substitution $\asc~=~\asu \cdot \rascasu$ and
$\asb =~\asu \cdot \rasbasu$. This substitution enables
\asu\ , \rascasu\ and \rasbasu\ to be determined
directly as free parameters in the fit and allows
correlations between these variables to be properly taken
into account.
In tests with Monte Carlo events it was verified
that this fitting procedure was sensitive to changes in \as\ for
c and b quarks with respect to uds quarks.
The fit ranges for each observable were determined by the range
of $y$ where the parton to hadron level corrections were below
10\% and the resulting $\chi^2/\mathrm{d.o.f.}$ remained
small ($\sim 5$ for $B_W$ and $\sim 2-3$ for the rest).
The remaining variations of the fit
results due to the choice of the fit ranges were taken as
systematic uncertainties. It was also checked that
within the fit range the \oaa\ calculation and the JETSET parton shower
model were in agreement with each other. The fit ranges used for this analysis
are slightly smaller than the fit ranges used in previous OPAL
studies~\cite{opalresummed, as_global} which used \oaa\ calculations in fits
to global event shapes. This is because the hadronization corrections for
distributions of event shape variables calculated with
b quark events are generally larger than the
corrections for distributions calculated with an inclusive event
sample. Consequently, the range of the distribution
in which the corrections are below 10\% is smaller.
The fit ranges, and
the results of the $\chi^2$ fits with $x_\mu = 1$ are shown in
Table~\ref{tab:oas2}.
The values for \asu\ obtained appear
large when compared with \as\ measurements performed using
resummed NLO calculations~\cite{opalresummed}.
This is due to the fact that fixed order
QCD calculations, valid only to \oaa , were used here.
To estimate the error in our results due to the choice of the
renormalization scale we adopted a method used in another
OPAL analysis~\cite{as_global} which used \oaa\ fits to global event
shapes. We performed fits with $x_\mu$ fixed equal one,
and fits in which $x_\mu$ was an additional free parameter. The average of
the two results is taken as our main result, and half
their difference as the ``scale uncertainty''.
The result of fits with $x_\mu$ as a free parameter are presented
in Table~\ref{tab:oas2free}. One sees that the values of
\asu\ decrease for all event shapes
and that the values of $\chi^2/$d.o.f. show a significant improvement.
This strong scale dependence is typical of determinations of \as\ where fixed
order calculations are used. It was checked that the values of \asu\
determined here were consistent with previous OPAL measurements of
\as~\cite{opalresummed,as_global}. The possibility of a different
scale dependence for heavy quark events, due to
mass effects present in higher order terms, was
investigated by introducing a separate $x_\mu$ parameter for each
quark flavour. The effect was found to be negligible within the
statistical precision of the measurement.
In Figs.~\ref{fig:d2th} through~\ref{fig:cp} the results of
fits with $x_\mu = 1$ are
plotted along with the measured distributions corrected to
the hadron level. The fit ranges used are indicated by the
arrow on each figure. It can be seen that within the fit ranges
the agreement between the fit results and the data are good.
In Fig.~\ref{fig:d2th} one sees a small disagreement
between the fit results and the data for values of $1-T > 0.25$.
When the fits to $1-T$ were repeated with $x_\mu$ as a free parameter,
the agreement between the fit results and data were
good over the entire fit range.
\section{Systematic uncertainties}
\label{sec:sys}
The main result was obtained
using the default selection and correction procedure described above.
The systematic uncertainties were divided into two groups:
experimental and theoretical uncertainties.
Each uncertainty was estimated by modifying details
of the event selection and correction procedure and
repeating the analysis. The difference between
the results obtained with the standard analysis and the
results obtained with the
analysis corresponding to each variation were taken (unless otherwise
noted) as symmetric systematic errors. In the case where a
parameter was varied above or below its nominal value, the
largest deviation from the main result was taken as a
symmetric systematic error associated with that parameter.
Finally, the systematic errors for each type
of variation were added in quadrature. The systematic
uncertainties investigated are described below and the effects they
had on the ratios \rascasu\ and \rasbasu\ are presented
in Tables~\ref{tab:sys_ascasu} and \ref{tab:sys_asbasu}.
\subsection{Experimental systematic uncertainties}
\begin{itemize}
\item The error due to the uncertainty on the flavour composition
of the flavour-tagged event samples was evaluated by varying the
flavour fractions in Table~\ref{tab:flavor} within their errors.
\item For the central result event shape observables
were measured using all
charged tracks and electromagnetic clusters.
To evaluate the relative response of the central tracking and
the electromagnetic calorimeter the measurements were performed
again using electromagnetic clusters only and charged tracks only.
\item The homogeneity of the response of the detector in the endcap region
was checked by restricting the analysis
to the barrel region of the
detector, requiring the thrust axis of accepted events to lie within
the range $|\cos\theta_{\rm Th}| < 0.7$.
This only modifies event shape distributions measured with
the c-tag event sample
because the uds-tag and b-tag event samples are already restricted
to the barrel region.
\item The minimum number of accepted charged tracks was
increased from 5 to 7 in order to further suppress background from
$\tau^+ \tau^-$ events and two-photon interactions.
\item The dependency of the result on the chosen fit range was
evaluated by adding or subtracting one bin from the lower and
upper end of the fit range. Each new fit range was treated as a
separate systematic variation.
\item In the charm event selection the requirement that
$x_{D^{*+}} > 0.4$ leads to a large kinematic bias in
the c-tag event sample. The uncertainty
due to this aspect of the $D^{*+}$ selection was evaluated by
varying the $x_{D^{*+}}$ cut between 0.3 and 0.5.
\end{itemize}
\subsection{Theoretical systematic uncertainties}
\begin{itemize}
\item The fragmentation of the final-state partons
into hadrons was modelled by JETSET
where the longitudinal fragmentation function for light
quarks was treated by the symmetric Lund model and
for heavy quarks by the model of Peterson et al~\cite{bib-PETERSON}.
The parameters for each model were determined from
a fit to OPAL data on global event shapes~\cite{bib-OPALtune}.
This fit yielded a value of $b = 0.52 \pm 0.04$ ({\tt PARJ(42)})
for the Lund model and $\epsilon_c = 0.031 \pm 0.011$
({\tt PARJ(54)}) and $\epsilon_b = 0.0038 \pm 0.0010$
({\tt PARJ(55)}) for the Peterson model.
To evaluate the systematic error
due to uncertainties in the JETSET hadronization model these
parameters were varied independently within one standard deviation of
their optimised values.
\item In JETSET, the parameter $Q_0$ ({\tt PARJ(82)}) is the virtuality cut-off
of partons. It determines the boundary between the
pertubative QCD and hadronization phases and is essentially
arbitrary. The optimum value to describe the OPAL data
was determined to be $Q_0 = (1.90 \pm 0.50)$\,GeV~\cite{bib-OPALtune} .
This parameter was varied within one standard deviation of its
optimised value.
\item The width of the transverse momentum distributions of quarks and antiquarks
produced in the fragmentation process is determined by the
parameter $\sigma_q = (0.40 \pm 0.03)$\,GeV ({\tt PARJ(21)}).
This parameter was varied within one standard deviation of its
optimum value.
\item The renormalization scale uncertainty was estimated by performing
fits with $x_\mu$ fixed equal one,
and fits in which $x_\mu$ is an additional free parameter. The average of
the two results is taken as our main result, and half
their difference as the scale uncertainty.
\item To evaluate the dependence of this analysis on the
choice of using JETSET to calculate the hadronization corrections,
the analysis was repeated using ARIADNE 4.08~\cite{ariadne} instead
of JETSET. The ARIADNE parton shower
is based upon a colour dipole model and provides an alternative to
the Lund parton shower model in JETSET. ARIADNE employs the same
fragmentation model as JETSET for the subsequent hadronization.
The difference between the results
using JETSET hadronization corrections and ARIADNE corrections were
used as an estimate of this systematic error.
\item The error due to uncertainties in the
c and b quark masses was evaluated by varying the
quark masses in the ranges given in~\cite{PDG} and repeating
the integration of the matrix elements corresponding to massive quarks
for each variation. The value for the b quark pole mass was
varied from 4.5 to 5.5\,GeV/$c^2$;
the c quark pole mass was varied from 1.2 to 1.9\,GeV/$c^2$.
\end{itemize}
\section{Results}
\label{sec:res}
The results of the $\chi^2$ fits to determine \rascasu\ and
\rasbasu\ for each of the event shape observables studied
are presented in Tables~\ref{tab:sys_ascasu} and \ref{tab:sys_asbasu}
and summarised in Fig.~\ref{fig:oas2}. The values quoted are
the average of fits with $x_\mu = 1$ and with $x_\mu$ as a free parameter.
The top half of the
Fig.~\ref{fig:oas2} shows the result of determining \rascasu\ and
the lower half \rasbasu . The vertical bars on the error bars
show the size of
the statistical error and the full error bar is the total
error, which is the sum of the
statistical, experimental systematic and theoretical errors
added in quadrature.
Also shown is the weighted mean of the results obtained
with the six observables. The weights were given by the
reciprocal of the square of the total error on
the ratios \rascasu\
and \rasbasu\ given in Tables~\ref{tab:sys_ascasu} and
\ref{tab:sys_asbasu}. The statistical error on the mean
was calculated taking into account correlations between
the five observables. The correlation
matrix was calculated from 100 Monte Carlo event samples.
The systematic uncertainty on the weighted mean was
determined from the change in the mean that occurred when each
systematic check was applied to all the event shape distributions
simultaneously.
The mean values of \rascasu\ and \rasbasu\
were determined to be
\begin{eqnarray}
\rascasu &=& 0.997
\pm 0.038 ~(stat.) \pm 0.030 ~(syst.) \pm 0.012 ~(theory) \nonumber \\
\rasbasu &=& 0.993
\pm 0.008 ~(stat.) \pm 0.006 ~(syst.) \pm 0.011 ~(theory)~. \nonumber
\end{eqnarray}
\noindent
In both cases the results are consistent with unity, indicating
flavour independence of \as .
As shown in Table~\ref{tab:sys_asbasu}, the largest experimental systematic error
on the results obtained from the variable $y_{23}$
originated from moving the lower bound of the fit range
from $y_{23} = 0.015$ to 0.025. It was found that this relatively large
variation in the result was due to a small discrepancy between the
data and the Monte Carlo, of the order of a few percent, for values of
$y_{23} < 0.03$. If one excludes the results obtained from $y_{23}$
from the calculation of the weighted means, the mean value obtained for
$\alpha_s^b/\alpha_s^{uds}$ remains unchanged within the experimental
precision, while the mean value of $\alpha_s^c/\alpha_s^{uds}$
changes from 0.997 to 0.990.
The size of the mass effect for c and b quarks
and its relevance to the measurement of \as\ are also of interest.
From phase space considerations one can estimate the size of the mass
effects for 3-jet observables. The ratio of the phase space of two massive
quarks and a gluon to the phase space for three massless particles is
$1 + 8(\rm{M}_q/\rm{M}_Z)^2 \rm{log}(\rm{M}_q/\rm{M}_Z)$~\cite{rodrigo_thesis}.
This represents a 7\% effect
for M$_q = 5$\,GeV/$c^2$ and 0.7\% for M$_q = 1.35$\,GeV/$c^2$.
Fig.~\ref{fig:masseff}a shows the
ratios \rascasu\ and \rasbasu\ determined using \oaa\ massless calculations
for both uds and c quarks and \oaa\ massive calculations for b quarks.
Again we show the average of fits with \xmu\ = 1.0 and \xmu\
as a free parameter.
One sees a small systematic shift of \rascasu\ with respect to the fits with
the massive calculation, although the
large statistical error on the measurements for c quarks makes a
definitive statement on the exact size of the mass effect difficult.
In Fig.~\ref{fig:masseff}b,
\rascasu\ and \rasbasu\ were determined using massless calculations for uds and
b quarks and massive calculations for c quarks.
In this case \rasbasu\ shows a large
systematic shift of order 5 to 7\%.
It should be noted that the values of
\asu\ and \rascasu\ remained unchanged within their statistical errors
with respect to the results presented in Fig.~\ref{fig:oas2}, therefore
the change in \rasbasu\ can be attributed entirely to changes in \asb .
This sensitivity to the b quark pole mass ( \mb\ ) suggests that this effect
could be exploited
to measure \mb\ itself, by assuming flavour independence of \as\ and
fitting for \mb . Monte Carlo studies show that the value of \mb\ determined
in this way depends strongly on the input value of \mb\ used in
JETSET to calculate hadronization corrections, thus making a determination
of \mb\ in this way problematic.
As expected, fits to $M_H$ do not exhibit
the mass effect like the other variables do. Here the quark mass
fixes a lower bound of the $M_H$ distribution since the invariant mass of the
jet cannot be less than the mass of the quark that originates the jet. For
\mb\ = 5.0\,GeV the lower bound lies at about $M_H/\sqrt{s} = 0.05$,
which is well outside of the fit range.
\section{Conclusion}
\label{sec:con}
We have presented a test of the flavour independence of the strong
coupling constant for charm and bottom quarks with respect to light
(uds) quarks. This analysis was based on a sample of hadronic decays
of the \znull\ resonance recorded by the OPAL detector at LEP.
The global event shapes $y_{23}$, $1-T$, $M_H$, $B_W$ and $C$
were used to measure \as\ in three flavour tagged event samples (uds, c and b).
The event shape distributions were fitted by \oaa\ calculations
of jet production
taking into account mass effects for the c and b quarks.
The ratios \rascasu\ and \rasbasu\ were both found to be
consistent with unity, indicating the flavour independence of \as .
The measurement of \rascasu\ achieved a precision of 5\%. The
relatively large statistical error was due to the low efficiency for
tagging c quark events ($\approx 2\%$) which relied on finding
D$^{*+}$ mesons.
The experimental error was dominated by varying the cut on the scaled
energy of the D$^{*+}$ mesons which was required to reduce the background
from c quark events coming from cascade decays of b hadrons.
The measurement of \rasbasu\ achieved a 1.5\% precision, the error being
dominated by the theoretical systematic error. The largest theoretical errors
were due to uncertainties in the b quark mass and
the renormalization scale factor $x_\mu$ .
In addition, we have presented a study of the effect of
heavy quark masses on global event shape variables.
It was observed that the values of
\as\ determined from $y_{23}$, $1-T$, $B_W$ and $C$
were reduced by 5 to 7\%
when these event shapes were measured
with a sample of b quark events and a massless QCD calculation was used.
The shape variable $M_H$ was found to be insensitive to the b quark mass.
\section{Acknowledgements}
\par
We particularly wish to thank the SL Division for the efficient operation
of the LEP accelerator at all energies
and for their continuing close cooperation with
our experimental group. We thank our colleagues from CEA, DAPNIA/SPP,
CE-Saclay for their efforts over the years on the time-of-flight and trigger
systems which we continue to use. In addition to the support staff at our own
institutions we are pleased to acknowledge the \\
Department of Energy, USA, \\
National Science Foundation, USA, \\
Particle Physics and Astronomy Research Council, UK, \\
Natural Sciences and Engineering Research Council, Canada, \\
Israel Science Foundation, administered by the Israel
Academy of Science and Humanities, \\
Minerva Gesellschaft, \\
Benoziyo Center for High Energy Physics,\\
Japanese Ministry of Education, Science and Culture (the
Monbusho) and a grant under the Monbusho International
Science Research Program,\\
Japanese Society for the Promotion of Science (JSPS),\\
German Israeli Bi-national Science Foundation (GIF), \\
Bundesministerium f\"ur Bildung, Wissenschaft,
Forschung und Technologie, Germany, \\
National Research Council of Canada, \\
Research Corporation, USA,\\
Hungarian Foundation for Scientific Research, OTKA T-016660,
T023793 and OTKA F-023259.\\
|
\section{Introduction}
Recent developments in the physics of granular matter \cite{jae-nag96}
have illustrated that the dissipative nature of the interactions
between grains can result in a variety of different phenomena. Of
particular interest in recent years has been the dynamics of vibrated
granular materials \cite{warretal95,meloetal95}, which exhibit
stationary states as well as waves and complex patterns. In order to
describe these diverse states of the material, it is necessary to
derive macroscopic descriptions by averaging over the microscopic
details of the motion and interactions between individual grains.
This goal has proved elusive, however, because a vibrated granular
material is a driven dissipative system, and the interactions between
the particles are characterised by a loss of energy due to inelastic
collisions. The statistical mechanics framework developed for
equilibrium or near equilibrium systems cannot be used in this case.
Consequently, phenomenological models
\cite{shrin97,tsim-aran97,venkat-ott98} have been used to describe the
dynamics of granular materials. The kinetic theories developed for
granular flows \cite{jen-sav83,kum98:vib} usually assume that the
system is close to ``equilibrium'' and the velocity distribution
function is close to the Maxwell-Boltzmann distribution.
Experimental studies and computer simulations have reported the
presence of a uniformly fluidised state in a vibrated bed of granular
material. Luding, Herrmann and Blumen \cite{ludetal94} carried out
`Event Driven' (ED) simulations of a two dimensional system of
inelastic disks in a gravitational field vibrated from below, and
obtained scaling laws for the density variations in the bed. An
experimental study of a vibrated fluidised bed was carried out by
Warr, Huntley and Jacques \cite{warretal95}. Their experimental set
up consisted of steel spheres confined between two glass plates that
are separated by a distance slightly larger than the diameter of the
spheres. The particles were fluidised by a vibrating surface at the
bottom of the bed, and the statistics of the velocity distribution of
the particles were obtained using visualisation techniques. Profiles
for the density and the mean square velocity were obtained, and the
particle velocity distributions were also determined at certain
positions in the bed. Both of these studies reported that there is an
exponential dependence of the density on the height near the top of
the bed, similar to the Boltzmann distribution for the density of a
gas in a gravitational field. However, the dependence of the density
deviates from the exponential behaviour near the bottom. The
dependence of the mean square velocity on the vibration frequency
and amplitude were found to be different in the two studies.
A theoretical calculation of the distribution function in a
vibro-fluidised bed was carried out by Kumaran
\cite{kum98:vib,kum98:vibscal}. The limit of low dissipation, where
the coefficient of restitution $e$ is close to $1$ was considered. In
this limit, the mean square velocity of the particles is large
compared to the mean square of the velocity of the vibrating surface,
and the dissipation of energy during a binary collision is small
compared to the energy of a particle. A perturbation approximation is
used, where the energy dissipation is neglected in the leading order
approximation, and the system resembles a gas at equilibrium in a
gravitational field. The velocity distribution function is a
Maxwell-Boltzmann distribution, and the density decreases
exponentially from the vibrating surface. The first order correction
to the distribution due to dissipative effects was calculated using
the moment expansion method, and the results were found to be in
qualitative agreement with the experiments of Warr et. al.
\cite{warretal95}.
The theoretical predictions \cite{kum98:vib,kum98:vibscal} were
compared with previous experimental and simulation studies by McNamara
and Luding \cite{mclud98}. They found that the theory was in good
agreement with experiments for dilute beds, where the area fraction of
the particles is low, but there were systematic deviations from the
theoretical predictions as the area fraction increases. This is to be
expected, since the analysis assumed that the density is small and the
pair distribution function was set equal to $1$ and therefore the
pressure is related to the density by the ideal gas law. These
assumptions become inaccurate as the area fraction of the bed
increases. An approximate method for including the correction to the
pair distribution function was suggested by Huntley \cite{hunt98}.
In the present analysis, the correction to the low density theory of
Kumaran \cite{kum98:vib,kum98:vibscal} is determined for a
vibro-fluidised bed where the coefficient of restitution is close to
$1$. An asymptotic analysis is used, where the dissipation is
neglected in the leading approximation. The leading order density and
velocity profiles are determined using the momentum balance equation
in the vertical direction. In contrast to the earlier theory
\cite{kum98:vib,kum98:vibscal}, the virial equation of state for a
non-ideal two dimensional gas is used to determine the leading order
density profile. The density profile differs from the Boltzmann
distribution, but the velocity distribution function is still a
Maxwell-Boltzmann distribution. The leading order temperature is
determined by a balance between the source and dissipation of energy
as before. The complete equilibrium pair distribution function is used
to determine the rate of dissipation of energy due to inelastic
collisions. The results are compared with hard sphere MD
simulations, and also with earlier theoretical and simulation studies.
\section{Analysis}
The system consists of a bed of circular disks (of diameter $\sigma$) in
a gravitational field driven by a vibrating surface. The vibrating
surface has a periodic amplitude function but no assumption is made
regarding the form of the function. There is a source of
energy at the vibrating surface due to particle collisions with the
surface, and the dissipation is due to inelastic collisions. A balance
between the two determines the ``temperature'', which is the mean
square velocity of the particles.
The limit of low dissipation, where the coefficient of restitution $e$
is close to $1$, is considered. In this limit, it can be shown that
the mean square velocity of the particles is large compared to the
mean square velocity of the vibrating surface. An asymptotic
expansion in the parameter $\epsilon \equiv U_0^2/T_0$ is used
\cite{kum98:vib}. If the source and dissipation of energy are
neglected in the leading approximation, the system resembles a gas of
hard disks at equilibrium in a gravitational field. The velocity
distribution function is a Maxwell-Boltzmann distribution at
equilibrium
\begin{equation}
F({\bf u}) = \frac{1}{2 \pi T_{0}} \exp{ \left( - \frac{u^{2}}{2
T_{0}} \right)},
\end{equation}
where $T_{0}$ is the leading order temperature. The density profile is
determined by solving the momentum balance equation in the vertical
direction
\begin{equation}
\label{eq:mombal}
\pder{p}{z} - \rho g = 0,
\end{equation}
where $p$ is the pressure, $\rho$ is the density (number of particles
per area) and $g$ is the acceleration due to gravity. For a gas at
equilibrium, the pressure is related to the density by the virial
equation of state, which in the case of inelastic circular disks is
\begin{equation}
p = \rho T_{0} \left[\frac{1 + e}{2} + (1 + e) g_0(\nu) \, \nu \right],
\end{equation}
where $g_{0}(\nu)$ is the pair distribution function at contact,
which for circular disks is given by \cite{verlet82}
\begin{equation}
g_{0}(\nu) = \frac{1}{16 (1 - \nu)^{2}} \left[ 16 - 7 \nu -
\frac{\nu^{3}}{4 (1 - \nu)^{2}} \right],
\end{equation}
and $\nu$ is the area fraction corresponding to $\rho$. If the
coefficient of restitution is set equal to $1$ in the leading
approximation, the equation for the pressure reduces to the standard
virial equation of state
\begin{equation}
p = \rho T_{0} \left[ 1 + 2 g_{0}(\nu) \, \nu\right].
\end{equation}
The resulting equation from \Eq{eq:mombal} for the density profile is
a first order ordinary differential equation, which can be solved
using the mass conservation condition
\begin{equation}
\label{eq:masscons}
\Int{0}{\infty}{z} \rho = N,
\end{equation}
where $N$ is the number of particles per unit width of the bed. Note
that the leading order temperature $T_{0}$ is still unknown at this
stage. This is determined using a balance between the source and
dissipation of energy. The source of energy due to particle
collisions with the vibrating surface is determined using an
equilibrium average over the increase in energy due to particle
collisions with the vibrating surface \cite{kum98:vib,kum98:vibscal}
\begin{equation}
\label{eq:s0}
S_{0} = 2 \sqrt{\frac{2}{\pi}} \, T_{0}^{1/2}
\braket{U^2} \, g_0(\nu) \, \rho \, \Big|_{z=0}.
\end{equation}
Here $\braket{U^2}$ represents the mean square velocity of the
vibrating surface. The rate of dissipation of energy per unit width
is calculated by averaging over the energy loss over all the
collisions between particles and integrating over the height of the
bed \cite{kum98:vib}
\begin{equation}
\label{eq:d0e}
D_0 = \sqrt{\pi} \, \sigma (1-e^2)\, T_0^{3/2}
\Int{0}{\infty}{z} g_0(\nu) \, \rho^2.
\end{equation}
Note that the $g_0$ appearing in $S_0$ and $D_0$ is the Enskog factor
which accounts for the increase in the frequency of collision for hard
disks at high densities. The temperature $T_{0}$ can now be
determined from the relation
\begin{equation}
\label{eq:s0d0}
S_0 = D_0
\end{equation}
An analytical solution to the density variation \Eq{eq:mombal} can be
determined in the low density limit using the equation of state for an
ideal gas for the pressure \cite{kum98:vib}.
\begin{equation}
\label{eq:nulow}
\rho = \frac{N g}{T_{0}} \exp{ \left( - \frac{g z}{T_{0}}
\right)}
\end{equation}
where the leading order temperature is given by,
\begin{equation}
T_{0} = \frac{4 \sqrt{2}}{\pi} \frac{\braket{U^2}}{N \sigma (1 -
e^{2})}.
\end{equation}
In the low-density limit the density decays exponentially from the
bottom of the bed. At higher densities the solution to the density
variation is no longer exponential throughout, and has to be obtained
numerically by an iterative scheme. However, at large distances from
the bottom, the bed is dilute and the ideal gas law holds good, hence
the decay is exponential, even though near the bottom it is not. This
gives a convenient starting point for the numerical integration from a
\emph{finite} height, above which we assume the asymptotic solution
($z \rightarrow \infty$) to be given by an exponential decay known to
within two undetermined constants. A value for the density and the
temperature is assumed at this height and the integration is carried
out up to the vibrating plate ($z=0$). The complete density profile
is obtained by combining the numerical and the asymptotic solutions.
If the conditions \Eq{eq:masscons} and \Eq{eq:s0d0} are not satisfied
after one such integration, a new value is determined for the density
and temperature using the Newton-Raphson method, and the iteration is
repeated till convergence. In cases where the convergence is poor, the
solution is obtained by \emph{continuing} a low density solution in a
parameter such as $N \sigma$ or $U_0$.
\textbf{Viscous dissipation:} The above analysis can be easily
extended to the case of dissipation purely due to viscous drag. The
expression for the source of energy remains the same as given by
\Eq{eq:s0}. A drag law given by $a_i = -\mu u_i$ is assumed. The total
leading order rate of dissipation per unit width will then be
\begin{eqnarray}
\label{eq:d0v}
D_{D0} & = & \mu \Int{0}{\infty}{z} \, \rho \Int{}{}{{\bf u}} \, F({\bf u})
\,\, {\bf u}\cdot{\bf u} \nonumber\\
& = & 2 \mu N T_0
\end{eqnarray}
Unlike \Eq{eq:d0e}, the leading order dissipation is the same for the
low density and the high density cases. Nevertheless, the density profile
has to be obtained numerically in the manner outlined above, with
\Eq{eq:d0v} substituted for \Eq{eq:d0e} in \Eq{eq:s0d0}.
\section{Simulation and Results}
The hard sphere molecular dynamics (MD), also known as event driven
(ED) method \cite{ludetal94} is used for the simulations of
the vibro-fluidised bed. Periodic boundary conditions are used in the
horizontal direction and the vibrating surface at the bottom has a
sawtooth form for the amplitude function. The simulations are carried
out only for the case of inelastic collisions, since the viscous drag
requires a different treatment than the ED method.
The density profiles obtained using the present analysis, as well as
the earlier low density approximations of Kumaran \cite{kum98:vib},
are compared with the simulation results in \Figs{fig:ldlenu}
and~\ref{fig:hdlenu}. It is seen that the density profiles of the
present analysis are in good agreement with the simulation results
even when the density near the bottom of the bed becomes large, while
the profiles from the low density approximation have significant
errors. \Fig{fig:hvis} shows the nature of the density profile in the
high density limit in the case of dissipation due to viscous drag.
Here too the present analysis gives reasonable values for packing
fraction near the bottom, while the low density theory predicts
physically incorrect values.
\begin{figure}[tb]
\begin{center}
\xlabel{$z/\sigma$}
\ylabel{$\nu$}
\gpfig{ldlenu.eps}
\caption{Exponential decay of packing fraction ($\nu$) with
a normalised height ($z/\sigma$) at low densities. The
predictions of the present analysis (solid line) and the low
density theory (dotted line) of \protect\cite{kum98:vib} is
compared with simulation (points). Both the predictions are
nearly identical. Here, $\epsilon=0.3$, $N \sigma=3$, $g=1$, and
$U_0=6$.}
\label{fig:ldlenu}
\end{center}
\end{figure}
\begin{figure}[tb]
\begin{center}
\xlabel{$z/\sigma$}
\ylabel{$\nu$}
\gpfig{hdlenu.eps}
\caption{Deviation of the density profile from the exponential decay
at high densities in the case of dissipation due to inelastic
collisions. The simulation result (points) is captured by the
present analysis (solid line) which is lower than the
exponential decay (dotted line) of the low density theory of
\protect\cite{kum98:vib} near the bottom of the bed. Here
$\epsilon=0.3$, $N \sigma=3$, $g=1$, and $U_0=1$.}
\label{fig:hdlenu}
\end{center}
\end{figure}
\begin{figure}[tb]
\begin{center}
\xlabel{$z/\sigma$}
\ylabel{$\nu$}
\gpfig{hvis.eps}
\caption{Deviation of the density profile from the exponential decay
at high densities in the case of dissipation due to viscous
drag. The present analysis (solid line) gives physically
plausible values for the packing fraction near the bottom, while
the low density theory (dotted line) of \protect\cite{kum98:vib}
predicts values higher than the maximum closed packing. Here
$\epsilon=0.2$, $N \sigma=20$, $g=20$, $\mu=0.1$, and
$U_0=5$.}
\label{fig:hvis}
\end{center}
\end{figure}
In a recent work, McNamara and Luding \cite{mclud98} reported the
scaling of dissipation with the center of mass obtained from
simulations. The results agreed with the low density theory of
\cite{kum98:vibscal} but a systematic deviation was observed at
high densities in all the cases. This deviation is captured in the
present analysis. The leading order dissipation at low densities in
the bed is given by \cite{kum98:vib}
\begin{equation}
\label{eq:d0}
D_0 = \frac{\sqrt{\pi}}{2}\, (1-e^2) \, N^2 \sigma g \sqrt{T_0}.
\end{equation}
In \cite{mclud98} the total dissipation obtained from the simulation
was normalised by a factor taken out from this leading order
dissipation and a non dimensional number was defined as
\begin{equation}
\label{e:cpp}
C_{pp} \equiv \frac{D_0}{(1-e) N^2 \sigma g \sqrt{T_0/2}}.
\end{equation}
The scaling of this factor with the height of the center of mass ($h$)
above the position at rest ($h_0$) was studied. This factor was found
out for different parameter sets by varying the bottom wall velocity
$U_0$ over several decades such that the bed is taken from a densely
packed regime to a very low density regime. They chose a central
data set and varied the parameters one at a time. It was found that
in all the cases considered, the scaling relation collapsed to a
single curve. The central parameter set has the following values $N =
3.2$, $\sigma=1$, $g=1$, $e=0.95$.
The present analysis is valid when $\epsilon \equiv U_0^2/T_0 \ll 1$
and when the frequency of particle-particle collision is much greater
than the frequency of particle-wall collisions. It can be shown that
in the leading order the ratio of the frequency of particle-particle
collisions to the frequency particle-wall collisions is $ \sqrt{2}\pi
\,N \sigma$. Hence the present analysis will hold good when $N \sigma \gg
1/\sqrt{2}\pi$. The central set corresponds to $\epsilon = 0.35$,
$N \sigma = 3.2$ and therefore we expect the present analysis to hold good
for this case. Most of the parameter sets used in \cite{mclud98} also
fall within the limits of the theory derived here.
\begin{figure}[tb]
\begin{center}
\xlabel{$2(h-h_0)/\sigma$}
\ylabel{$C_{pp}$}
\gpfig{theo.eps}
\caption{Theoretical scaling of the normalised dissipation
($C_{pp}$) against the center of mass ($h$) above the position
at rest ($h_0$) for the different cases reported in
\protect\cite{mclud98}. All except two---(N+) with $\epsilon =
1.73$ and (N--) with $N \sigma = 0.65$ collapse on to a single curve
in the linear region. The parameters indicated correspond to
$N=16 $ (N+), $N=0.65$ (N--), $g=25$ (g+),
$g=0.04$ (g--), $e=0.99$ (e+), $e=0.75$ (e--), rest of the
parameters being same as the one in the central set, which has
the following values $N = 3.2$, $\sigma=1$, $g=1$,
$e=0.95$.}
\label{fig:theoscal}
\end{center}
\end{figure}
\begin{figure}[tb]
\begin{center}
\xlabel{$2(h-h_0)/\sigma$}
\ylabel{$C_{pp}$}
\gpfig{simtheo.eps}
\caption{Scaling of the normalised dissipation with the center
of mass: Predictions from the present analysis is compared with
the results from our simulations and the reported results in
\protect\cite{mclud98}. The linear portion of all the curves
from theory, except two, fall on the solid line denoted as
`Theory'. The two exceptions are also shown. A set of points
correspond to the simulation data with parameter values $N =
16$ (N+), $N=0.65$ (N--), $g=25$ (g+),
$g=0.04$ (g--), $e=0.99$ (e+), $e=0.75$ (e--); rest of the
parameters in a set being the same as the one in the central
set, which has the following values $N = 3.2$, $\sigma=1$,
$g=1$, $e=0.95$.}
\label{fig:simtheo}
\end{center}
\end{figure}
\Fig{fig:theoscal} shows the theoretical predictions of the total
dissipation for the different cases reported in Fig.~2 in
\cite{mclud98}. It is compared with the results of two simulations in
\Fig{fig:simtheo}. It is seen that the present analysis correctly
predicts the lowering of the coefficient $C_{pp}$ at high densities.
This reduction in the dissipation from the constant value at low
densities is the net result of two opposing factors: (i) decrease in
the density from the exponential behaviour near the vibrating bottom
(see \Fig{fig:hdlenu}), hence reducing the total value of the
dissipation, and (ii) increase in frequency of collisions at high
densities, increasing the dissipation.
It is also seen that not all the theoretical predictions collapse on
to a curve as is the case with the data from the simulation. In two of
the cases the theory does not agree with the simulations because (i)
in one the value of the perturbation parameter is high ($\epsilon =
1.73$) and the leading order theory is valid only for low $\epsilon$,
and (ii) in the other case the value of $N \sigma = 0.65 $ is low.
In \Fig{fig:theoscal}, the apparent mismatch with `e-' is not a
discrepancy with the model, but has got to do with the formula chosen
used in \cite{mclud98} for the normalisation of the dissipation factor
$C_{pp}$. They had chosen to normalise the dissipation by a factor
$(1-e)$. While this might have given a better fit for high densities
(low center of mass), the correct factor for very low densities is
$(1-e^2)$ as given by \Eq{eq:d0}. The difference is more pronounced in
the case of $e\ll 1$, which, here, has a value $e=0.75$. A close
inspection of the curves `e-' in \Fig{fig:theoscal} and
\Fig{fig:simtheo} show that the theory and simulation do indeed agree
with each other.
We also note here that the data taken from the reported simulation
\cite{mclud98} is for asymmetric sawtooth vibration, whereas our
simulation is for the symmetric sawtooth. Both these give similar
results for the scaling of $C_{pp}$. Also the theoretical predictions
for the symmetric and the asymmetric sawtooth are identical,
indicating that the form of the bottom wall vibration does not affect
the scaling of the dissipation with the center of mass.
\section{Conclusion}
In summary, a theory to describe the state of a vibro-fluidised bed in
the dense limit was derived. This is different from the earlier theory
of Kumaran \cite{kum98:vib,kum98:vibscal}, which is valid for low
densities where the ideal gas equation was used and the pair
distribution function was set equal to $1$. We have made use of the
virial equation of state to obtain a correction to the exponential
density profile obtained in low densities and the pair distribution
function is used to calculate the increased frequency of collisions in
the source and the dissipation of energy. The theoretical predictions
of density and temperature were compared with the results obtained
from MD simulation of two dimensional disks. The theory correctly
predicts the lowering of the density from the exponential value at
high densities near the bottom. The theory also predicts the scaling
relations of the total dissipation in the bed reported in
\cite{mclud98}.
|
\section{Introduction}
How many perfect matchings does a given graph $G$ have?
That is, in how many ways can one choose a subset of the edges of $G$
so that each vertex of $G$ belongs to one and only one chosen edge?
(See Figure 1(a) for an example of a perfect matching of a graph.)
For general graphs $G$,
it is computationally hard to obtain the answer \cite{PrV},
and even when we have the answer, it is not so clear
that we are any the wiser for knowing this number.
However, for
many infinite families of special graphs the number of perfect matchings
is given by compellingly simple formulas.
Over the past ten years a great many families of this kind
have been discovered,
and while there is no single unified result that encompasses all of them,
many of these families resemble one another,
both in terms of the form of the results
and in terms of the methods that have been useful in proving them.
\begin{figure}
\centerline{\psfig{file=\figdir/azgraph,scale=110}}
\caption{The Aztec diamond of order 4.}
\label{azgraph}
\end{figure}
The deeper significance of these formulas is not clear.
Some of them are related to results in representation theory
or the theory of symmetric functions,
but others seem to be self-contained combinatorial puzzles.
Much of the motivation for this branch of research
lies in the fact that we are still unable to predict ahead of time
which enumerative problems lead to beautiful formulas and which do not;
each new positive result seems like an undeserved windfall.
Hereafter, I will use the term ``matching'' to signify ``perfect matching''.
(See the book of Lov\'asz and Plummer \citeyear{PrLP}
for general background on the theory of matchings.)
As far as I have been able to determine, problems involving enumeration
of matchings were first examined by chemists and physicists in the 1930s,
for two different (and unrelated) purposes: the study of aromatic
hydrocarbons and the attempt to create a theory of the liquid state.
Shortly after the advent of quantum chemistry, chemists turned their
attentions to molecules like benzene composed of carbon rings with
attached hydrogen atoms. For these researchers, matchings of a graph
corresponded to ``Kekul\'e
structures'', i.e., ways of assigning single and double bonds in the
associated hydrocarbon (with carbon atoms at the vertices and tacit
hydrogen atoms attached to carbon atoms with only two neighboring
carbon atoms). See for example the article of Gordon and Davison
\citeyear{PrGD},
whose use of nonintersecting lattice paths anticipates certain
later work \cite{PrGV,PrSac1,PrJS}. There
are strong connections between combinatorics and chemistry for such
molecules; for instance, those edges which are present in comparatively
few of the matchings of a graph turn out to correspond to the bonds that
are least stable, and the more matchings a polyhex graph possesses the
more stable is the corresponding benzenoid molecule. Since hexagonal
rings are so predominant in the structure of hydrocarbons, chemists gave
most of their attention to counting matchings of subgraphs of the infinite
honeycomb grid.
At approximately the same time, scientists were trying to understand
the behavior of liquids. As an extension of a more basic model for
liquids containing only molecules of one type, Fowler and Rushbrooke
\citeyear{PrFR} devised a lattice-based model for liquids containing two
types of molecules, one large and one small. In the case where the
large molecule was roughly twice the size of the small molecule, it
made sense to model the small molecules as occupying sites of a
three-dimensional grid and the large molecules as occupying pairs of
adjacent sites. In modern parlance, this is a monomer-dimer model.
In later years, the two-dimensional version of the model was found
to have applicability to the study of molecules adsorbed on films;
if the adsorption sites are assumed to form a lattice, and an adsorbed
molecule is assumed to occupy two such sites, then one can imagine
fictitious molecules that occupy all the unoccupied sites (one each).
Major progress was made when Temperley and Fisher \citeyear{PrTF}
and Kasteleyn \citeyear{PrKa1} independently found ways to count pure
dimer configurations on subgraphs of the infinite square grid,
with no monomers present. Although the physical significance of
this special case was (and remains) unclear, this result, along with
Onsager's earlier exact solution of the two-dimensional Ising model
\cite{PrO}, paved the way for other advances such as Lieb's exact solution
of the six-vertex model \cite{PrL}, culminating in a new field at the
intersection of physics and mathematics: exactly solved
statistical mechanics models in two-dimensional lattices. (Intriguingly,
virtually none of the three- and higher-dimensional analogues of these
models have succumbed to researchers' efforts at obtaining exact solutions.)
For background on lattice models in statistical mechanics, see
the book by Baxter \citeyear{PrB}.
An infinite two-dimensional grid has many finite subgraphs; in choosing
which ones to study, physicists were guided by
the idea that the shape of boundary should be chosen so as to minimize
the effect of the boundary\emdash that is, to maximize the number of
configurations, at least in the asymptotic sense. For example,
Kasteleyn, in his study of the dimer model on the square grid,
counted the matchings of the $m$-by-$n$ rectangle (see the double-product
formula at the beginning of Section~5) and of the $m$-by-$n$ rectangular torus,
and showed that the two numbers grow at the
same rate as $m,n$ go to infinity, namely $C^{mn}$ for a known constant $C$.
(Analytically, $C$ is $e^{G/\pi}$, where $G$ is Catalan's constant
$1 - \frac{1}{9} + \frac{1}{25} - \frac{1}{49} + \frac{1}{81} - \cdots{}$;
numerically, $C$ is approximately $1.34$.)
Kasteleyn \citeyear{PrKa1} wrote: ``The effect of boundary
conditions is, however, not
entirely trivial and will be discussed in more detail in a subsequent
paper.'' (See the article of Cohn, Kenyon and Propp \cite{PrCKP}
for a rigorous mathematical treatment of boundary conditions.)
Kasteleyn never wrote such a followup paper, but other physicists did
give some attention to the issue of boundary shape, most notably
Grensing, Carlsen and Zapp \cite{PrGCZ}. These authors
considered a one-parameter family of graphs of the kind
shown in Figure 1(a), and they asserted that every graph in this family
has $2^{N/4}$ matchings, where $N$
is the number of vertices. They did not give a proof, nor did they indicate
whether they had one. The result was rediscovered in the late 1980s by
Elkies, Kuperberg, Larsen, and Propp \cite{PrEKLP}, who gave four proofs of
the formula. This article led to a great deal of work among enumerative
combinatorialists, who refer to graphs like the one shown in Figure 1
as ``Aztec diamond graphs'', or sometimes just Aztec diamonds for short.
(It should be noted that Elkies et al.\ \citeyear{PrEKLP} used the term ``Aztec
diamond'' to denote regions like the one shown in Figure 1(b). The two
sorts of Aztec diamonds are dual to one another; matchings of Aztec diamond
graphs correspond to domino tilings of Aztec diamond regions.)
At about the same time, it became clear that there had been earlier work
within the combinatorial community that was pertinent to the study of
matchings, though its relevance had not hitherto been recognized. For
instance, Mills, Robbins and Rumsey \cite{PrMRR}, in their work on alternating
sign matrices, had counted pairs of ``compatible'' ASMs of consecutive size;
these can be put into one-to-one correspondence with matchings
of an associated Aztec diamond graph \cite{PrEKLP}.
Looking into earlier mathematical literature,
one can even see intimations of enumerative matching theory in
the work of MacMahon \citeyear{PrM}, who nearly a century ago found a formula
for the number of plane partitions whose solid Young diagram fits inside
an $a$-by-$b$-by-$c$ box, as will be discussed in Section~2.
(See the book by Andrews
\citeyear{PrA} and the article by Stanley \citeyear{PrSt1} for
background on plane partitions.)
Such a Young diagram is nothing more than an assemblage of cubes,
and it has long been known in the extra-mathematical world
that such assemblages, viewed from a distant point, looks like tilings
(consider Islamic art, for instance). Thus it was natural for
mathematicians to interpret MacMahon's theorem on plane partitions
as a result about tilings of a hexagon by rhombuses.
This insight may have occurred to a number of people independently;
the earliest chain of oral communication that I have followed
leads back to Klarner (who did not publish his observation
but relayed it to Stanley in the 1970s),
and the earliest published statement I have found is in a paper
by David and Tomei \citeyear{PrDT}.{\looseness=-1\par}
In any case, each of the Young diagrams enumerated by MacMahon
corresponds to a tiling of a hexagon by rhombuses,
where the hexagon is semiregular
(its opposite sides are parallel and of equal length,
with all internal angles equal to 120 degrees)
and has side-lengths $a,b,c,a,b,c$,
and where the rhombuses have all side-lengths equal to 1.
These tilings in turn correspond to matchings of the ``honeycomb''
graph that is dual to the dissection of the hexagon into unit equilateral
triangles; see Figure 2, which shows a matching of the honeycomb graph
and the associated tiling of a hexagon.
Kuperberg \citeyear{PrKu1} was the first to exploit the connection
between plane partitions and the dimer model.
(Interestingly, some of the same graphs
that Kuperberg studied had been investigated independently by chemists
in their study of benzenoids hydrocarbons; Cyvin and Gutman \citeyear{PrCG}
give a survey of this work.)
\begin{figure}
\centerline{\psfig{file=\figdir/hexgraph,scale=110}}
\caption{A matching and its associated tiling.}
\label{hexgraph}
\end{figure}
Similarly, variants of MacMahon's problem in which the plane
partition is subjected to various symmetry constraints
(considered by Macdonald, Stanley, and others
[\citeNP{PrSt3}; \citeyearNP{PrSt4}])
correspond to the problem of enumerating matchings possessing
corresponding kinds of symmetry. Kuperberg \citeyear{PrKu1} used
this correspondence in solving one of Stanley's open problems, and this
created further interest in matchings among combinatorialists.
One of Kuperberg's chief tools was an old result of Kasteleyn, which
showed that for any planar graph $G$, the number of matchings
of $G$ is equal to the Pfaffian of a certain matrix of zeros and ones
associated with $G$. A special case of this result, enunciated by
Percus \citeyear{PrPe},
can be used when $G$ is bipartite; in this case, one can use a determinant
instead of a Pfaffian. Percus' determinant is a modified version of
the bipartite adjacency matrix of the graph, in which rows correspond to
``white'' vertices and columns correspond to ``black'' vertices (under
a coloring scheme whereby white vertices have only black neighbors and
vice versa); the $(i,j)$-th entry is $\pm 1$ if the $i$-th white vertex and
$j$-th black vertex are adjacent, and 0 otherwise. For more details on
how the signs of the entries are chosen, see the
expositions of Kasteleyn \citeyear{PrKa3} and Percus \citeyear{PrPe}.
Percus' theorem, incorporated into computer software, makes it easy to
count the matchings of many planar graphs and look for patterns in
the numbers that arise. Two such programs are {\tt vaxmaple}, written
by Greg Kuperberg, David Wilson and myself, and {\tt vaxmacs}, written
by David Wilson. Most of the patterns described below were
discovered with the aid of this software, which is available from
\url{http://math.wisc.edu/~propp/software.html}. Both programs
treat subgraphs of the infinite square grid; this might seem restrictive,
but it turns out that counting the matchings of an arbitrary bipartite
planar graph can be fit into this framework, with a bit of tweaking.
The mathematically interesting part of each program is the routine for
choosing the signs of the nonzero entries. There are many choices
that would work, but Wilson's sign-rule is far and away the simplest:
If an edge is horizontal, we give it weight $+1$, and if an edge is
vertical, joining a vertex in one row to a vertex in the row below it,
we give the edge weight $(-1)^k$, where $k$ is the number of vertices in
the upper row to the left of the vertical edge.
The main difference between {\tt vaxmaple} and {\tt vaxmacs} is that
the former creates \Maple/ code which, if sent to \Maple/,
results in \Maple/ printing out the number of matchings of the graph;
{\tt vaxmacs}, on the other hand, is a customized {\sc Emacs} environment
that fully integrates text-editing operations (used for defining the
graph one wishes to study) with the mathematical operations of interest.
Both programs represent bipartite planar graphs in ``VAX-format'',
where V's, A's, X's, and other letters denote vertices. (An example
of VAX-format can be found on page~\pageref{vaxexample}; for a detailed
explanation see \url{http://math.wisc.edu/~propp/vaxmaple.doc}.)
Quite recently, the study of matchings of nonbipartite graphs
has been expedited by the programs {\tt graph} and {\tt planemaple},
created by Matt Blum and Ben Wieland, respectively.
These programs make it easy to define a planar graph by pointing and clicking,
after which one can count its matchings using an efficient implementation
of Kasteleyn's Pfaffian method. This makes it easy to try out new ideas
and look for patterns, outside of the better-explored bipartite
case.{\looseness=1\par}
Interested readers with access to the World Wide Web
can obtain copies of all of these programs via
\url{http://math.wisc.edu/~propp/software.html}.
Most of the formulas that have been discovered
express the number of matchings of a graph
as a product of many comparatively small factors.
Even before one has conjectured (let alone proved) such a formula,
one can frequently infer its existence
from the fact that the number of matchings
has only small prime factors.
Numbers that are large compared to their largest prime factor
are sometimes called ``smooth'' or ``round'';
the latter term will be used here.
The definition of roundness is not precise,
since it is not intended for use as a technical term.
Its vagueness is intended to capture the uncertainties and the suspense
of formula-hunting,
and the debatable issue of whether the occurrence of a single
larger-than-expected prime factor
rules out the existence of a product formula.
(For an example of a number whose roundness lies in this gray area,
see the table of numbers given in Problem 8.)
It is worth noting that Kuperberg \citeyear[Section VII-A]{PrKu2}
has shown that rigorous proofs of roundness need not always yield explicit
product formulas.
Christian Krattenthaler has written a Mathematica
program called RATE that greatly expedites the process of guessing patterns
in experimental data on enumeration of matchings; see
\url{http://radon.mat.univie.ac.at/People/kratt/rate/rate.html}.
A great source of the appeal of research on enumeration of matchings is the
ease with which undergraduate research assistants can participate in the hunt
for formulas and proofs; many members of the M.I.T.\ Tilings Research Group
(composed mostly of undergraduates like Blum and Wieland) played a role in
the developments that led to the writing of this article. Enumeration of
matchings has turned out to be a rich avenue of combinatorial inquiry,
and many more beautiful patterns undoubtedly await discovery.
Updates on the status of these problems can be found on the Web at
\url{http://math.wisc.edu/~propp/update.ps.gz}.
\section{Lozenges}
We begin with problems related to lozenge tilings of hexagons.
A \textit{lozenge} is a rhombus of side-length 1
whose internal angles measure 60 and 120 degrees;
all the hexagons we will consider will tacitly have
integer side-lengths and internal angles of 120 degrees.
Every such hexagon $H$ can be dissected into unit equilateral triangles
in a unique way,
and one can use this dissection to define a graph $G$
whose vertices correspond to the triangles
and whose edges correspond to pairs of triangles that share an edge;
this is the ``finite honeycomb graph''
dual to the dissection.
It is easy to see
that the tilings of $H$ by lozenges are in one-to-one correspondence
with the matchings of $G$.
The $a,b,c$ semiregular hexagon
is the hexagon whose side lengths are, in cyclical order, $a,b,c,a,b,c$.
Lozenge tilings of this region
are in correspondence with
plane partitions with
at most $a$ rows,
at most $b$ columns,
and no part exceeding $c$.
Such hexagons are represented in VAX-format by diagrams like
$$
\label{vaxexample}
\vcenter\brda{%
AVAVAVAVA
AVAVAVAVAVA
AVAVAVAVAVAVA
AVAVAVAVAVAVAVA
VAVAVAVAVAVAVAV
VAVAVAVAVAVAV
VAVAVAVAVAV
VAVAVAVAV
}
$$
where A's and V's represent upward-pointing and downward-pointing triangles,
respectively. In this article we will use triangles instead:
$$\vcenter\brd{
AVAVAVAVA
AVAVAVAVAVA
AVAVAVAVAVAVA
AVAVAVAVAVAVAVA
VAVAVAVAVAVAVAV
VAVAVAVAVAVAV
VAVAVAVAVAV
VAVAVAVAV
}
$$
MacMahon \citeyear{PrM} showed that the number of such plane partitions is
$$\prod_{i=0}^{a-1} \prod_{j=0}^{b-1} \prod_{k=0}^{c-1}
\frac{i+j+k+2}{i+j+k+1}\,.$$
(This form of MacMahon's formula is due to Macdonald; a short,
self-contained proof is given by Cohn et al. \citeyear[Section~2]{PrCLP}.)
\begin{problem}
\label{prob1}
Show that in the $2n-1$, $2n$, $2n-1$ semiregular hexagon,
the central location (consisting of the two innermost triangles) is
covered by a lozenge in exactly one-third of the tilings.
(Equivalently: Show that if one chooses a random matching of
the dual graph, the probability that the central edge is contained
in the matching is exactly $\frac13$.)
\end{problem}
\begin{progress}
Two independent and very different solutions of this problem have been
found; one by Mihai Ciucu and Christian Krattenthaler
and the other by Harald Helfgott and Ira Gessel.
Ciucu and Krattenthaler \citeyear{PrCK}
compute more generally the number of rhombus tilings of a hexagon
with sides $a,a,b,a,a,b$ that contain the central unit rhombus, where $a$
and $b$ must have opposite parity (the special case $a=2n-1$, $b=2n$ solves
Problem 1). The same generalization was obtained (in a different but
equivalent form) by Helfgott and Gessel \citeyear{PrHG}, using a completely
different method.
One might still try to look for a proof whose simplicity
is comparable to that of the answer ``one-third''.
Also worthy of note is the paper of Fulmek and Krattenthaler \citeyear{PrFK1},
which generalizes the result of Ciucu and Krattenthaler \citeyear{PrCK}.
\end{progress}
\hskip\parindent
The hexagon of side-lengths
$n$, $n+1$, $n$, $n+1$, $n$, $n+1$ cannot be tiled by
lozenges at all, for in the dissection into unit triangles, the number of
upward-pointing triangles differs from the number of downward-pointing
triangles. However, if one removes the central triangle, one gets a
region that can be tiled, and the sort of numbers one gets for small
values of $n$ are striking. Here they are, in factored form:
$$
\medmuskip 1mu
\displaylines{
2\cr
2\cdot3^3\cr
2^5\cdot3^3\cdot5\cr
2^5\cdot5^7\cr
2^2\cdot5^7\cdot7^5\cr
2^8\cdot3^3\cdot5\cdot7^{11}\cr
2^{13}\cdot3^9\cdot7^{11}\cdot11\cr
2^{13}\cdot3^{18}\cdot7^5\cdot11^7\cr
2^8\cdot3^{18}\cdot11^{13}\cdot13^5\cr
2^2\cdot3^9\cdot11^{19}\cdot13^{11}\cr
2^{10}\cdot3^3\cdot11^{19}\cdot13^{17}\cdot17\cr
2^{16}\cdot11^{13}\cdot13^{23}\cdot17^7\cr
}
$$
These are similar to the numbers one gets from counting lozenge tilings
of an $n,n,n,n,n,n$ hexagon, in that the largest prime factor seems to
be bounded by a linear function of $n$.
\begin{problem}
\label{prob2}
Enumerate the lozenge tilings of the region obtained
from the $n$, $n+1 $, $n $, $n+1 $, $n $, $n+1$ hexagon by removing
the central triangle.
\end{problem}
\begin{progress}
Mihai Ciucu has solved the more general problem of counting the rhombus
tilings of an $(a$, $b+1 $, $b $, $a+1 $, $b $, $b+1)$-hexagon with the central triangle removed
\cite{PrCi2}.
Ira Gessel proved this result independently using the nonintersecting
lattice-paths method \cite{PrHG}.
Soichi Okada and Christian Krattenthaler have solved the even more general
problem of counting the rhombus tilings of an $(a $, $b+1 $, $c $, $a+1 $, $b $, $c+1)$-hexagon
with the central triangle removed \cite{PrOK}.
\end{progress}
\hskip\parindent
One can also take a $2n $, $2n+3 $, $2n $, $2n+3 $, $2n $, $2n+3$ hexagon
and make it lozenge-tilable by removing a triangle from the
middle of each of its three long sides, as shown:
$$\vcenter\brd{
AVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVAVAVAVA
VAVAVAVAVAVAVAVAVAVAVAVAV
AVAVAVAVAVAVAVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVA
AVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVA
VAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAV
VAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAV
VAVAVAVAVAVAVAVAVAVAVAVAVAVAVAVAV
VAVAVAVAVAVAVAVAVAVAVAVAVAVAVAV
VAVAVAVAVAVAVAVAVAVAVAVAVAVAV
VAVAVAVAVAVAVAVAVAVAVAVAVAV
VAVAVAVAVAVAVAVAVAVAVAVAV
VAVAVAVAVAV VAVAVAVAVAV
}$$
Here one obtains an equally tantalizing sequence of factorizations:
$$
\medmuskip 1mu
\displaylines{
1\cr
2^7\cdot 7^2\cr
2^2\cdot 7^4\cdot 11^4\cdot 13^2\cr
2^{10}\cdot 3^3\cdot 5^8\cdot 13^2\cdot 17^4 \cdot 19^2\cr
2^2\cdot 5^2\cdot 7^2\cdot 11^3\cdot 13^4\cdot 17^4\cdot 19^8\cdot 23^4
}
$$
\begin{problem}
\label{prob3}
Enumerate the lozenge tilings of the region obtained
from the $2n$, $2n+3 $, $2n $, $2n+3 $, $2n $, $2n+3$ hexagon by
removing a triangle from
the middle of each of its long sides.
\end{problem}
\begin{progress}
Theresia Eisenk\"olbl solved this problem. What she does in fact is
to compute the number of all rhombus tilings of a hexagon with sides
$a$, $b+3$, $c$, $a+3$, $b$, $c+3$, where an arbitrary triangle is
removed from each of the ``long'' sides of the hexagon (not
necessarily the triangle in the middle). For the proof of her formula
\cite{PrE1} she uses nonintersecting lattice paths, determinants, and
the Jacobi determinant formula \cite{PrT}. However, I still know of no
conceptual explanation for why these numbers are so close (in the
multiplicative sense) to being perfect squares.
\end{progress}
\hskip\parindent
We now return to ordinary $a,b,c$ semiregular hexagons.
When $a=b=c$, there are not two but six central triangles.
There are two geometrically distinct ways in which we can choose to
remove an upward-pointing triangle and downward-pointing triangle
from these six, according to whether the triangles are opposite or adjacent:
$$\vcenter\brd{
AVAVAVA AVAVAVA
AVAVAVAVA AVAVAVAVA
AVAVA AVAVA AVAV VAVAVA
VAVAV VAVAV VAVA AVAVAV
VAVAVAVAV VAVAVAVAV
VAVAVAV VAVAVAV
}
\vadjust{\vskip3pt}
$$
Such regions may be called ``holey hexagons'' of two different kinds.
Matt Blum tabulated the number of lozenge tilings of these regions,
for small values of $a=b=c$.
In the first (``opposite'') case, the number
of tilings of the holey hexagon is a nice round number (its greatest
prime factor appears to be bounded by a linear function of the size
of the region). In the second (``adjacent'') case, the number
of tilings is not round. Note, however, that in the second case,
the number of tilings of the holey hexagon divided by the number of tilings
of the unaltered hexagon (given to us by MacMahon's formula) is equal to
the probability that a random lozenge tiling of the hexagon contains a
lozenge that covers these two triangles; this probability tends to $\frac13$
for large $a$, at least on average \cite{PrCLP}.
Following this clue, we examine the difference between the aforementioned
probability (with its messy, un-round numerator) and the number $\frac13$.
The result is a fraction in which the numerator is now a nice round number.
So, in both cases, we have reason to think that there is an exact product
formula.
\begin{problem}
\label{prob4}
Determine the number of lozenge tilings of a
regular hexagon from which two of its innermost unit triangles
(one upward-pointing and one downward-pointing) have been removed.
\end{problem}
\begin{progress}
Theresia Eisenk\"olbl solved the first case of Problem 4
and Markus Fulmek and Christian Krattenthaler solved the second case.
Eisenk\"olbl \citeyear{PrE2} solves a generalization of the problem by applying
Mihai Ciucu's matchings factorization theorem, nonintersecting lattice
paths, and a nontrivial determinant evaluation. Fulmek and Krattenthaler
\citeyear{PrFK2} compute the number of rhombus tilings of a hexagon
with sides $a,b,a,a,b,a$ (with $a$ and $b$ having the same parity)
that contain the rhombus that touches the center of the hexagon and
lies symmetric with respect to the symmetry axis that runs parallel
to the sides of length $b$.
For the proof of their formula they compute Hankel determinants
featuring Bernoulli numbers, which they do by using facts about continued
fractions, orthogonal polynomials, and, in particular, continuous Hahn
polynomials. The special case $a=b$ solves the second part of Problem 4.
\end{progress}
\hskip\parindent
I mentioned earlier that Kasteleyn's method, as interpreted by Percus,
allows one to write the number of matchings of a bipartite planar graph as
the determinant of a signed version of the bipartite adjacency matrix.
In the case of lozenge tilings of hexagons and the associated matchings,
it turns out that there is no need to modify signs of entries; the ordinary
bipartite adjacency matrix will do. Greg Kuperberg \citeyear{PrKu2}
has noticed that when row-reduction and column-reduction are
systematically applied to the Kasteleyn--Percus matrix of
an $a,b,c$ semiregular hexagon,
one can obtain the $b$-by-$b$ Carlitz matrix \cite{PrCS}
whose $(i,j)$-th entry is
$a+c \choose a+i-j$.
(This matrix can also be recognized as the Gessel--Viennot matrix
that arises from interpreting each tiling
as a family of nonintersecting lattice paths \cite{PrGV}.)
Such reductions do not affect the determinant,
so we have a pleasing way of understanding the relationship
between the Kasteleyn--Percus matrix method
and the Gessel--Viennot lattice-path method.
In fact, such reductions
do not affect the \textit{cokernel} of the matrix
(an abelian group whose order is the determinant).
On the other hand, the cokernel of the Kasteleyn--Percus matrix
for the $a,b,c$ hexagon is clearly invariant under
permuting $a$, $b$, and $c$.
This gives rise to three different Carlitz matrices
that nontrivially have the same cokernel.
For example,
if $c=1$,
then one gets an $a$-by-$a$ matrix and a $b$-by-$b$ matrix
that both have the same cokernel,
whose structure can be determined ``by inspection'' if one notices
that the third Carlitz matrix of the trio is just a 1-by-1 matrix
whose sole entry is (plus or minus) a binomial coefficient.
In this special case,
the cokernel is just a cyclic group.
Greg Kuperberg poses this challenge:
\begin{problem}
\label{prob5}
Determine the cokernel of the Carlitz matrix,
or equivalently of the Kasteleyn--Percus matrix of the $a,b,c$ hexagon,
and if possible find a way to interpret the cokernel in terms of
the tilings.
This combines Questions 1 and 2 of Kuperberg \citeyear{PrKu2}.
As he points out in that article, in the case $a=b=c=2$,
one gets the noncyclic group $\Z/2\Z \times \Z/10\Z$ as the cokernel.
\end{problem}
\hskip\parindent
As was remarked above, one nice thing about the Kasteleyn--Percus matrices
of honeycomb graphs is that it is not necessary to make any of the
entries negative. For general graphs,
however, there is no canonical way of defining $K$, in the sense
that there may be many ways of modifying the signs of certain
entries of the bipartite adjacency matrix of a graph so that all
nonzero contributions to the determinant have the same sign.
Thus, one should not expect the eigenvalues of $K$ to possess
combinatorial significance. However, the spectrum of $K$ times
its adjoint $K^*$ is independent of which Kasteleyn--Percus matrix $K$
one chooses (as was independently shown by David Wilson and Horst
Sachs). Thus, digressing somewhat from the topic of lozenge tilings,
we find it natural to ask:
\begin{problem}
\label{prob6}
What is the significance of the spectrum of
$K K^*$, where $K$ is any Kasteleyn--Percus matrix associated with a
bipartite planar graph?
\end{problem}
\begin{progress}
Nicolau Saldanha \citeyear{PrSal} has proposed a combinatorial
interpretation of the spectrum of $K K^*$.
Horst Sachs says (personal communication)
that $K K^*$ may have some significance
in the chemistry of polycyclic hydrocarbons (so-called
benzenoids) and related compounds as a useful approximate
measure of the ``degree of aromaticity''.
\end{progress}
\hskip\parindent
Returning now to lozenge tilings, or equivalently, matchings
of finite subgraphs of the infinite honeycomb,
consider the hexagon graph with $a=b=c=2$:%
$$
\vcenter{\rot{\psfig{file=\figdir/hex,width=1.4in}}{0}{0}}
$$
This is the graph whose 20 matchings correspond to the 20 tilings
of the regular hexagon of side 2 by rhombuses of side 1. If we look
at the probability of each individual vertical edge belonging to a
matching chosen uniformly at random (``edge-probabilities''), we get
\vadjust{\vskip4pt}
$$
\makeatletter
\vcenter{\rot{\psfig{file=\figdir/hexa,width=1.4in}}{0}{0}}
$$
Now look at this table of numbers as if it described a distribution
of mass. If we assign the three rows $y$-coordinates $-1$ through 1,
we find that the weighted sum of the squares of the $y$-coordinates is equal to
$$(0.3+0.4+0.3)(-1)^2+(0.7+0.3+0.7+0.3)(0)^2+(0.3+0.4+0.3)(1)^2=2.$$
If we assign to the seven
columns $x$-coordinates $-3$ through 3, we find that the weighted sum of the
squares of the $x$-coordinates is equal to
$(0.7)(-3)^2+(0.6)(-2)^2+(0.3)(-1)^2+(0.8)(0)^2
+(0.3)(1)^2+(0.6)(2)^2+(0.7)(3)^2=20$.
\vadjust{\goodbreak}%
You can do a similar (but even easier) calculation yourself for the case
$a=b=c=1$, to see that the ``moments of inertia'' of the vertical
edge-probabilities around the horizontal and vertical axes are 0 and
1, respectively. Using {\tt vaxmaple} to study the case $a=b=c=n$ for
larger values of $n$, I find that the moment of inertia about the horizontal
axis goes like
$$0, 2, 12, 40, 100, \dots$$
and the moment of inertia about the vertical axis goes like
$$1, 20, 93, 296, 725, \dots .$$
It is easy to show that the former moments of inertia
are given in general by the polynomial
$(n^4-n^2)/6$
(in fact, the number of vertical lozenges that have any particular
$y$-coordinate does not depend on the tiling chosen).
The latter moments of inertia are subtler;
they are not given by a polynomial of degree 4,
though it is noteworthy that the $n$-th term
is an integer divisible by $n$,
at least for the first few values of $n$.
\begin{problem}
\label{prob7}
Find the ``moments of inertia'' for the mass on
edges arising from edge-probabilities for random matchings of the
$a,b,c$ honeycomb graph.
\end{problem}
\section{Dominoes}
Now let us turn from lozenge-tiling problems to domino-tiling problems.
A \textit{domino} is a 1-by-2 or 2-by-1 rectangle.
Although lozenge tilings (in the guise of constrained plane partitions)
were studied first, it was really the study of domino tilings in Aztec
diamonds that gave current work on enumeration of matchings its current
impetus. Here is the Aztec diamond of order 5:
$$\vcenter\brd{
XX
XXXX
XXXXXX
XXXXXXXX
XXXXXXXXXX
XXXXXXXXXX
XXXXXXXX
XXXXXX
XXXX
XX
}$$
A tiling of such a region by dominos is equivalent to a matching
of a certain (dual) subgraph of the infinite square graph.
This grid is bipartite, and it is convenient to color its vertices
alternately black and white;
equivalently, it is convenient to color the 1-by-1 squares alternately
black and white, so that every domino contains
one 1-by-1 square of each color.
Elkies, Kuperberg, Larsen, and Propp showed
in \cite{PrEKLP}
that the number of domino tilings
of such a region is $2^{n(n+1)/2}$ (where $2n$ is the number of rows),
and Gessel, Ionescu, and Propp proved in \cite{PrGIP} an exact formula
(originally conjectured by Jockusch) for the number of tilings of regions like
$$\vcenter\brd{
XX
XXXX
XXXXXX
XXXXXXXX
XXXX XXXXX
XXXX XXXXX
XXXXXXXX
XXXXXX
XXXX
XX
}$$
in which two innermost squares of opposite color have been removed.
(For some values of $n$, the number of tilings is exactly $\frac14$
times $2^{n(n+1)/2}$; in the other cases, there is an exact product
formula for the difference between the number of tilings and
$\bigl(\frac14\bigr)2^{n(n+1)/2}$. It is this latter fact that motivated the
idea of trying something similar in the case of lozenge tilings,
as described in the paragraph preceding the statement of Problem 4.)
Now suppose one removes two squares from the middle of an Aztec diamond of
order $n$ in the following way:
$$\vcenter\brd{
XX
XXXX
XXXXXX
XXXX XXX
XXXXXXXXXX
XXXX XXXXX
XXXXXXXX
XXXXXX
XXXX
XX
}$$
(The two squares removed are a knight's-move apart, and subject to that
constraint, they are as close to being in the middle as they can be.
Up to symmetries of the square, there is only one way of doing this.)
The numbers of tilings one gets are as follows (for $n = 2$ through 10):
$$
\medmuskip 1mu
\displaylines{
2\cr
2^3\cr
2^5\cdot 5\cr
2^9\cdot 3^2\cr
2^{17}\cdot 3\cr
2^{22}\cdot 3^2\cr
2^{24}\cdot 3^2\cdot73\cr
2^{31}\cdot 3^2\cdot5^2\cdot11\cr
2^{47}\cdot 3^2\cdot5\cr
}
$$
Only the presence of the large prime factor 73 makes one doubt
that there is a general product formula; the other prime factors are
reassuringly small.
\begin{problem}
\label{prob8}
Count the domino tilings of an Aztec diamond from which
two close-to-central squares, related by a knight's move, have been
deleted.
\end{problem}
\begin{progress}
Harald Helfgott has solved this problem;
it follows from the main result in his thesis \citeyear{PrH}.
The formula is somewhat complicated,
as the prime factor 73 might have led us to expect.
(One of the factors in Helfgott's product formula
is a single-indexed sum;
73 arises as $128-60+5$.)
\end{progress}
\hskip\parindent
One can also look at ``Aztec rectangles'' from which squares have been
removed so as to restore the balance between black and white squares
(a necessary condition for tileability). For instance, one can remove
the central square from an $a$-by-$b$ Aztec rectangle in which $a$
and $b$ differ by 1, with the larger of $a,b$ odd:
$$\vcenter\brd{
XX
XXXX
XXXXXX
XXXXXXXX
XXXX XXXX
XXXXXXXX
XXXXXX
XXXX
XX
}$$
\begin{problem}
\label{prob9}
Find a formula for the number of domino tilings of a
$2n$-by-$(2n+1)$ Aztec rectangle with its central square removed.
\end{problem}
\begin{progress}
This had already been solved when I posed the problem; it is
a special case of a result of Ciucu \citeyear[Theorem 4.1]{PrCi1}.
Eric Kuo solved the problem independently.
\end{progress}
\hskip\parindent
What about $(2n-1)$-by-$2n$ rectangles? For these regions, removing
the central square does not make the region tilable. However, if
one removes any one of the four squares adjacent to the middle square,
one obtains a region that is tilable, and moreover, for this region
the number of tilings appears to be a nice round number.
\begin{problem}
\label{prob10}
Find a formula for the number of domino tilings of a
$(2n-1)$-by-$2n$ Aztec rectangle with a square adjoining the central
square removed.
\end{problem}
\begin{progress}
This problem was solved independently three times:\,\
by Harald Helfgott and Ira Gessel \citeyear{PrHG},
by Christian Krattenthaler \citeyear{PrKr},
and by Eric Kuo (private communication).
Gessel and Helfgott solve a more general problem than Problem 10.
Krattenthaler's preprint
gives several results concerning the enumeration of matchings
of Aztec rectangles where (a suitable number of) collinear vertices are
removed, of which Problem 10 is just a special case.
There is some overlap between the results of Helfgott and Gessel
and the results of Krattenthaler.
\end{progress}
\hskip\parindent
At this point, some readers may be
wondering why $m$-by-$n$ rectangles have not played
a bigger part in the story.
Indeed, one of the surprising facts of life in the study of enumeration
of matchings is that Aztec diamonds and their kin have been
much more fertile ground for exact combinatorics that the seemingly
more natural rectangles. There are, however,
a few cases I know of in which something rather nice turns up. One
is the problem of Ira Gessel that appears as Problem 20 in this document.
Another is the work done by Jockusch \citeyear{PrJ} and, later, Ciucu \citeyear{PrCi1}
on why the number
of domino tilings of the square is always either a perfect square or
twice a perfect square. In the spirit of the work of Jockusch and
Ciucu, I offer here a problem based on Lior Pachter's observation
\cite{PrPK} that the region on the left below, obtained by removing
8 dominos from a 16-by-16 square, has exactly one tiling.
What if we make the intrusion half as long, as in the region on the right?
$$\vcenter\brd{
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXX XXXXXXX
XXXXXX XXXXXXXX
XXXXX XXXXXXXXX
XXXX XXXXXXXXXX
XXX XXXXXXXXXXX
XX XXXXXXXXXXXX
X XXXXXXXXXXXXX
XXXXXXXXXXXXXX
}\hskip .25\hsize
\vcenter\brd{
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXX XXXXXXXXXXX
XX XXXXXXXXXXXX
X XXXXXXXXXXXXX
XXXXXXXXXXXXXX
}$$
That is, we take a $2n$-by-$2n$ square (with $n$ even) and remove $n/2$
dominos from it, in a partial zig-zag pattern that starts from the corner.
Here are the numbers we get, in factored form, for $n=2,4,6,8,10$:
$$
\medmuskip 1mu
\displaylines{
2\cdot3^2\cr
2^2\cdot3^6\cdot13^2\cr
2^3\cdot3^2\cdot5^4\cdot7^2\cdot3187^2\cr
2^4\cdot11771899^2\cdot27487^2\cr
2^5\cdot2534588575976069659^2
}$$
The factors are ugly, but the exponents are nice: we get $2^{n/2}$ times
an odd square.
Perhaps this is a special case of a two-parameter fact that says that
you can take an intrusion of length $m$ in a $2n$-by-$2n$ square and the
number of tilings of the resulting region will always be a square or
twice a square.
\begin{problem}
\label{prob11}
What is going on with ``intruded Aztec diamonds''?
In particular, why is the number of tilings so square-ish?
\end{problem}
\hskip\parindent
It should also be noted that the square root of the odd parts of these
numbers (3, $3^3 \cdot 13$, etc.)\ alternate between 1 and 3 mod 4.
Perhaps these quantities are continuous functions of $n$ in the 2-adic
sense, as is the case for intact $2n$-by-$2n$ squares \cite{PrCo};
however, the presence of large prime factors means that no simple
product formula is available, and that the analysis will require new
techniques.
\medbreak
We now return to the Kasteleyn--Percus matrices discussed earlier.
Work of Rick Kenyon and David Wilson \cite{PrKe}
has shown that the \textit{inverses}
of these matrices are loaded with combinatorial information, so it
would be nice to get our hands on them. Unfortunately, there are
many nonzero entries in the inverse-matrices. (Recall that
the Kasteleyn--Percus matrices themselves, being nothing more than adjacency
matrices in which some of the 1's have been strategically replaced
by $-1$'s, are sparse; their inverses, however, tend to have most
if not all of their entries nonzero.) Nonetheless, some exploratory
``numerology'' leaves room for hope that this is do-able.
Consider the Kasteleyn--Percus matrix $K_n$ for the Aztec diamond of
order $n$, in which every vertical domino with its white square on top
(relative to some fixed checkerboard coloring) has its sign
inverted\emdash that is, the corresponding 1 in the bipartite
adjacency matrix is replaced by $-1$.
\begin{problem}
\label{prob12}
Show that the sum of the entries of the matrix inverse
of $K_n$ is $\frac12(n-1)(n+3) - 2^{n-1} + 2$.
\end{problem}
\hskip\parindent
(This formula works for $n=1$ through $n=8$.)
\begin{progress}
Harald Helfgott has solved a similar problem using the main result of
his thesis
\citeyear{PrH}, and it is likely that the result asserted in Problem 12
can be proved similarly.
(A slight technical hurdle arises from the fact that Helfgott's thesis uses
a different sign-convention for the Kasteleyn--Percus matrix, which
results in different signs, and a different sum, for the inverse
matrix; however, Helfgott's methods are quite general, so there is
no conceptual obstacle to applying them to Problem 12.)
I should mention that my original reason for examining the
sum of the entries of the inverse Kasteleyn--Percus matrix
was to see whether there might be formulas governing the
individual entries themselves. Helfgott's work provides
such formulas.
Also, in this connection, Greg Kuperberg and Douglas
Zare have some high-tech ruminations on the inverses of Kasteleyn--Percus
matrices, and there is a chance that representation-theory methods will
give a different way of proving the result.
\end{progress}
\hskip\parindent
Now we turn to a class of regions I call ``pillows''.
Here are a ``0 mod 4'' pillow of ``order 5'' and a
``2 mod 4'' pillow of ``order 7'':
$$\vcenter\brd{
XXXX
XXXXXXXX
XXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXX
XXXXXXXXXXXX
XXXXXXXX
XXXX
}\hskip.1\hsize
\vcenter\brd{
XX
XXXXXX
XXXXXXXXXX
XXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXXXXXX
XXXXXXXXXXXXXX
XXXXXXXXXX
XXXXXX
XX
}$$
It turns out (empirically) that the number of tilings of the 0-mod-4 pillow
of order $n$ is a perfect square times the coefficient of $x^n$ in the Taylor
expansion of $(5+3x+x^2-x^3)/(1-2x-2x^2-2x^3+x^4)$. This fact came to light
in several steps. First it was noticed that the number of tilings has a
comparatively small square-free part. Then it was noticed that in the
derived sequence of square-free parts, many terms were roughly three times
the preceding term. Then it was noticed that, by
judiciously including some of the square factors, one could obtain
a sequence in which each term was roughly three times the preceding term.
Finally it was noticed that this approximately geometric sequence satisfied
a fourth-order linear recurrence relation.
Similarly, it appears that the number of tilings of the 2-mod-4 pillow of
order $n$ is a perfect square times the coefficient of $x^n$ in the Taylor
expansion of $(5+6x+3x^2-2x^3)/(1-2x-2x^2-2x^3+x^4)$. (If you are wondering
about ``odd pillows'', I should mention that there is a nice formula for the
number of tilings, but this is not an interesting result, because an odd pillow
splits up into many small noncommunicating sub-regions such that a tiling
of the whole region corresponds to a choice of tiling on each of the
sub-regions.)
\begin{problem}
\label{prob13}
Find a general formula for the number of domino tilings
of even pillows.
\end{problem}
\hskip\parindent
Jockusch looked at the Aztec diamond of order $n$ with a 2-by-2
hole in the center, for small values of $n$;
he came up with a conjecture for the number of domino tilings,
subsequently proved by Gessel, Ionescu, and Propp \cite{PrGIP}.
One way to generalize this is to make the hole larger,
as was suggested by Douglas Zare
and investigated by David Wilson.
Here is an abridged and adapted version of the report
David Wilson sent me on October 15, 1996:
\medskip
Define the Aztec window with outer order $y$ and
inner order $x$ to be the Aztec diamond of order $y$ with an
Aztec diamond of order $x$ deleted from its center. For
example, this is the Aztec window with orders 8 and 2:
$$\vcenter\brd{
XX
XXXX
XXXXXX
XXXXXXXX
XXXXXXXXXX
XXXXXXXXXXXX
XXXXXX XXXXXX
XXXXXX XXXXXX
XXXXXX XXXXXX
XXXXXX XXXXXX
XXXXXXXXXXXX
XXXXXXXXXX
XXXXXXXX
XXXXXX
XXXX
XX
}$$
There are a number of interesting patterns that show up
when we count tilings of Aztec windows. For one thing,
if $w$ is a fixed even number, and $y = x+w$, then for any $w$
the number of tilings appears to be a polynomial in $x$.
(When $w$ is odd, and $x$ is large enough, there are no tilings.)
For $w=6$, the polynomial is
$$
\displaylines{
8192 x^8 + 98304 x^7 + 573440 x^6 + 2064384 x^5 + 4988928 x^4 \hfill\cr
\hfill + 8257536 x^3 + 9175040 x^2 + 6291456 x + 2097152. }
$$
This can be written as
$$
2^{17} \bigl( \tfrac{1}{2} \bigl(x+\tfrac{3}{2}\bigr)^2 + \tfrac{7}{8} \bigr)^4
$$
or as
$$
2^{17}x^4 \ \ \circ \ \
\tfrac{1}{2}x+\tfrac{7}{8} \ \ \circ \ \ \bigl(x+\tfrac{3}{2}\bigr)^2,
$$
where it is understood that these three polynomials get composed.
More generally, all the polynomials in $x$ that arise in this fashion
appear to ``factor'' in the sense of functional composition. Here
are the factored forms of the polynomials for $n=2,4,6,8,10$:
$$
\arraycolsep=4pt
\begin{array}{rcccl}
2^{3}x^4&\circ&1&\circ&\bigl(x+\frac12\bigr)^2\\[2ex]
2^{8}x^2&\circ&x+1&\circ&\bigl(x+1\bigr)^2\\[2ex]
2^{17}x^4&\circ&\frac12x+\frac{7}{8}&\circ&\bigl(x+\frac32\bigr)^2\\[2ex]
2^{28}x^2&\circ&\frac{1}{144}x^4+\frac{7}{72}x^3
+\frac{41}{144}x^2+\frac{11}{18}x+1&\circ&\bigl(x+2\bigr)^2\\[2ex]
2^{43}x^4&\circ&\frac{1}{144}x^3+\frac{61}{576}x^2
+\frac{451}{2304}x+\frac{967}{1024}&\circ&\bigl(x+\frac52\bigr)^2
\end{array}
$$
In general the rightmost polynomial is $(x+w/4)^2$,
and the leftmost polynomial is either a perfect square,
twice a fourth power, or half a fourth power, depending
on $w$ mod 8. A pattern for the middle polynomial however
is elusive.
\begin{problem}
\label{prob14}
Find a general formula for the number of domino tilings
of Aztec windows.
\end{problem}
\begin{progress}
Constantin Chiscanu
found a polynomial bound on
the number of domino tilings of the Aztec window
of inner order $x$ and outer order $x+w$ \cite{PrCh}.
Douglas Zare used the transfer-matrix method to show that
the number of tilings is not just bounded by a polynomial,
but given by a polynomial, for each fixed $w$ \cite{PrZ}.
\end{progress}
\section{Miscellaneous}
Now we come to some problems involving tiling that fit neither the
domino-tiling nor the lozenge-tiling framework.
Here the more general picture is that we have some periodic dissection
of the plane by polygons,
such that an even number of polygons meet at each vertex,
allowing us to color the polygons alternately black or white.
We then make a suitable choice of a finite region $R$
composed of equal numbers of black and white polygons,
and we look at the number of ``diform'' tilings of the region,
where a \textit{diform} is the union of two polygonal cells
that share an edge.
In the case of domino tilings, the underlying dissection of the infinite plane
is the tiling by squares, 4 around each vertex;
in the case of lozenge tilings, the underlying dissection of the infinite plane
is the tiling by equilateral triangles, 6 around each vertex.
Other sorts of periodic dissections have already played a role in
the theory of enumeration of matchings.
For instance, there is a tiling of the plane by isosceles right triangles
associated with a discrete reflection group in the plane;
in this case, the right choice of $R$ (see Figure 3)
gives us a region that can be tiled in $5^{n^2/4}$ ways when $n$ is even
and in $5^{(n^2-1)/2}$ or $2 \cdot 5^{(n^2-1)/2}$ ways when $n$ is odd
\cite{PrY}.
\begin{figure}
\vskip1pt
\centerline{\psfig{file=\figdir/fortress}}
\vskip-1pt
\caption{A fortress of order 5, with $2 \times 5^6$ diform tilings.}
\label{fortress}
\end{figure}
\begin{figure}[b]
\vskip1pt
\centerline{\psfig{file=\figdir/azdungeon}}
\vskip-1pt
\caption{An Aztec dungeon of order 2, with $13^3$ diform tilings.}
\label{azdungeon}
\end{figure}
Similarly, in the tiling of the plane by triangles
that comes from a 30 degree, 60 degree, 90 degree right triangle
by repeatedly reflecting it in its edges,
a certain region called the ``Aztec dungeon''
(see Figure 4)
gives rise to a tiling problem in which powers of 13 occur
(as was proved in not-yet-published work of Mihai Ciucu).
A key feature of these regions $R$
is revealed by looking at the colors of those polygons in the dissection
that share an edge with the border of $R$.
One sees that the border splits up into four long stretches
such that along each stretch,
all the polygons that touch the border have the same color.
It is not clear why regions with this sort of property
should be the ones that give rise to
the nicest enumerations,
but this appears to happen in practice.
One interesting case
arises from a rather symmetric
dissection of the plane into
equilateral triangles, squares, and regular hexagons,
with 4 polygons meeting at each vertex
and with no two squares sharing an edge.
A typical diform tiling of this region
(called a ``dragon'')
is shown in Figure 5.
\begin{figure}
\centerline{\psfig{file=\figdir/dragon}}
\caption{A dragon of order 10 (tiled).}
\label{trisquex}
\end{figure}
Empirically, one finds that the number of diform tilings is $2^{n(n+1)}$.
\begin{problem}
\label{prob15}
Prove that the number of diform
tilings of the dragon of order $n$ is $2^{n(n+1)}$.
\end{problem}
\begin{progress}
Ben Wieland solved this problem (private communication).
\end{progress}
\hskip\parindent
Incidentally, the tiling shown in Figure 5 was generated
using an algorithm that generates each of the possible diform tilings
of the region with equal probability.
It is no fluke that the tiling looks so orderly
in the left and right corners of the region;
this appears to be typical behavior in situations of this kind.
This phenomenon has been analyzed rigorously for two tiling-models:
lozenge tilings of hexagons \cite{PrCLP} and domino tilings of Aztec diamonds
\cite{PrCEP}.
One way to get a new dissection of the plane from an old one is to refine it.
For instance, starting from the dissection of the plane into squares,
one can draw in every $k$-th southwest-to-northeast diagonal. When
$k$ is 1, this is just a distortion of the dissection of the plane into
equilateral triangles. When $k$ is 2, this is a dissection that leads to
finite regions for which the number of diform tilings is a known power of 2,
thanks to a theorem of Chris Douglas \citeyear{PrD}.
But what about $k=3$ and higher?
For instance, we have the roughly hexagonal region shown in Figure 6;
certain boundary vertices have been marked with a dot so as to
bring out the large-scale $2,3,2,2,3,2$ hexagonal structure more clearly.
\begin{figure}[b]
\centerline{\psfig{file=\figdir/wieland,width=2in}}
\caption{A region for Problem 16.}
\label{wieland}
\end{figure}
The cells of this region are triangles and squares.
The region has $17920 = 2^9 \cdot 5 \cdot 7$ diform tilings.
\begin{problem}
\label{prob16}
Find a formula for the number of diform tilings in the
$a,b,c$ quasihexagon in the dissection of the plane that arises from
slicing the dissection into squares along every third upward-sloping
diagonal.
\end{problem}
\hskip\parindent
One reason for my special interest in Problem 16
is that it seems to be a genuine hybrid of domino tilings of Aztec diamonds
and lozenge tilings of hexagons.
\begin{progress}
Ben Wieland solved this problem in the case $a=b=c$ (which, as it
turns out, is also the solution to the case $a=b<c$ and the case
$a=c<b$). In these cases the number of tilings is always a power
of two. The general case does not yield round numbers, so there
is no simple product formula.
\end{progress}
\hskip\parindent
The approach underlying Ben Wieland's solutions to the last two problems
is a method of subgraph substitution that has already been of
great use in enumeration of matchings of graphs.
I will not go into great detail here on this method
[\citeNP{PrPr1}; \citeyearNP{PrPr2}],
but here is an overview:
One studies graphs with weights assigned to their edges,
and one does weighted enumeration of matchings,
where the weight of a matching is the product
of the weights of the constituent edges.
One then looks at local substitutions of subgraphs within a graph that
preserve the sum of the weights of the matchings,
or more generally,
multiply the sum of the weights of the matchings
by some predictable factor.
Then the problem of weight-enumerating matchings of one graph
reduces to the problem of weight-enumerating matchings
of another graph.
Iterating this procedure,
one can often eventually reduce the graph
to something easier to understand.
\medskip
Problems 15 and 16 are just two instances of a broad class of problems
arising from periodic graphs in the plane. A unified understanding
of this class of problems has begun to emerge, by way of subgraph
substitution. The most important open problem connected with this class
of results is the following:
\begin{problem}
\label{prob17}
Characterize those local substitutions that have
a predictable effect on the weighted sum of matchings of a graph.
\end{problem}
\begin{figure}[b]
\centerline{\psfig{file=\figdir/urban,width=3in}}
\caption{The ``urban renewal'' substitution.}
\label{urban}
\end{figure}
\hskip\parindent
The most useful local substitution so far has been the one shown in
Figure~7, where unmarked edges have weight 1 and where $A,B,C,D$ are
respectively obtained from $a,b,c,d$ by dividing by $ad+bc$; if $G$
and $G'$ denote the graph before and after the substitution, one can
check that the sum of the weights of the matchings of $G'$ equals the
sum of the weights of the matchings of $G$ divided by $ad+bc$.
It is required that the four innermost vertices
have no neighbors other than the four vertices shown;
this constraint is indicated by circling them.
Noncircled vertices may have any number of neighbors.
\begin{figure}[h]
\vskip2pt
\centerline{\psfig{file=\figdir/kenyon,width=2in}}
\vskip-2pt
\caption{Rick Kenyon's substitution.}
\label{kenyon}
\end{figure}
The substitution shown in Figure 8
(a straightforward generalization of
a clever substitution due to Rick Kenyon)
has also been of use.
Here the new weights are not entirely determined by the old,
but have a single degree of freedom;
the relevant formulas can be written as
$$
\def\hskip 10 pt minus 6pt{\hskip 10 pt minus 6pt}
A = \frac{abc+aeg+cdf}{bc+eg} \,, \hskip 10 pt minus 6pt B = b \,, \hskip 10 pt minus 6pt
D = \frac{dg}{bc+eg} E \,, \hskip 10 pt minus 6pt
F = ef \frac{1}{E} \,,\hskip 10 pt minus 6pt
G = (bc+eg) \frac{1}{E}\,,
$$
with $E$ free.
As before, the circled vertices
must not have any neighbors
other than the ones shown.
In this case, the sum of the weights in the before-graph $G$
is exactly equal to the sum of the weights in the after-graph $G'$;
there is no need for a correction factor
like the $1/(ad+bc)$ that arises in urban renewal.
The extremely powerful ``wye-delta'' substitution of
Colbourn, Provan, and Vertigan \cite{PrCPV} should also be mentioned.
\medskip
Up till now we have been dealing exclusively with bipartite planar graphs.
We now turn to the less well-explored nonbipartite case.
For instance, one can look at the triangle graph of order $n$,
shown in Figure~9 in the case $n=4$.
(Here $n$ is the number of vertices in the longest row.)
\begin{figure}
\centerline{\psfig{file=\figdir/triangle,width=1.1in}}
\caption{The triangle graph.}
\label{triangle}
\end{figure}
Let $M(n)$ denote the number of matchings of the triangle graph of order $n$.
When $n$ is 1 or 2 mod 4, the graph has an odd number of vertices
and $M(n)$ is 0; hence let us only consider the cases in which
$n$ is 0 or 3 mod 4.
Here are the
first few values of $M(n)$, expressed in factored form:
$2$,
$2 \cdot 3$,
$2 \cdot 2 \cdot 3 \cdot 3 \cdot 61$,
$2 \cdot 2 \cdot 11 \cdot 29 \cdot 29$,
$2^3 \cdot 3^3 \cdot 5^2 \cdot 7^2 \cdot 19 \cdot 461$,
$2^3 \cdot 5^2 \cdot 37^2 \cdot 41 \cdot 139^2$,
$2^4 \cdot 73 \cdot 149 \cdot 757 \cdot 33721 \cdot 523657$,
$2^4 \cdot 3^8 \cdot 17 \cdot 37^2 \cdot 703459^2$,
\dots.
It is interesting that $M(n)$ seems to be divisible by
$2^{\lfloor (n+1)/4 \rfloor}$
but no higher power of 2;
it is also interesting that when we divide by this power of 2,
in the case where $n$ is a multiple of 4,
the quotient we get, in addition to being odd,
is a perfect square times a small number
$(3, 11, 41, 17, \dots{})$.
\begin{problem}
\label{prob18}
How many matchings does the triangle graph
of order $n$ have?
\end{problem}
\begin{progress}
Horst Sachs \citeyear{PrSac2} has responded to this problem.
\end{progress}
\hskip\parindent
One can also look at graphs that are bipartite but not planar.
A natural example is the $n$-cube (that is, the $n$-dimensional cube
with $2^n$ vertices). It has been shown that the number
of matchings of the $n$-cube goes like $1$, $2$, $9=3^2$,
$272=16 \cdot 17$,
$589185=3^2 \cdot 5 \cdot 13093$, \dots.
\begin{problem}
\label{prob19}
Find a formula for the number of matchings
of the $n$-cube.
\end{problem}
\hskip\parindent
(This may be intractable; after all, the graph has exponentially
many vertices.)
\begin{progress}
L\'aszl\'o Lov\'asz gave a simple proof of my (oral) conjecture that
the number of matchings of the $n$-cube
has the same parity as $n$ itself.
Consider the orbit of a particular matching of the $n$-cube
under the group generated by the $n$ standard reflections of the $n$-cube.
If all the edges are parallel (which can happen in exactly $n$ ways),
the orbit has size 1;
otherwise the size of the orbit is
of the form $2^k$ (with $k \geq 1$)\emdash an even number.
The claim follows,
and similar albeit more complex reasoning
should allow one to compute the enumerating sequence
modulo any power of 2.
Meanwhile,
L. H.\ Clark, J. C.\ George, and T. D.\ Porter have shown \cite{PrCGP}
that if one lets $f(n)$ denote the number of 1-factors in the $n$-cube,
then $$f(n)^{2^{1-n}} \sim n/e$$
as $n \rightarrow \infty$.
It was subsequently pointed out by Bruce Sagan
that the main result of Clark et al.\ \citeyear{PrCGP} is a special case
of the theorem cited by Lov\'asz and Plummer \citeyear[top of page 312]{PrLP}.
\end{progress}
\hskip\parindent
Finally, we turn to a problem involving domino tilings of rectangles,
submitted by Ira Gessel (what follows are his words):
\catcode `!=11
\newdimen\squaresize
\newdimen\thickness
\newdimen\Thickness
\newdimen\ll! \newdimen \uu! \newdimen\dd!
\newdimen \rr! \newdimen \temp!
\def\sq!#1#2#3#4#5{%
\ll!=#1 \uu!=#2 \dd!=#3 \rr!=#4
\setbox0=\hbox{%
\temp!=\squaresize\advance\temp! by .5\uu!
\rlap{\kern -.5\ll!
\vbox{\hrule height \temp! width#1 depth .5\dd!}}%
\temp!=\squaresize\advance\temp! by -.5\uu!
\rlap{\raise\temp!
\vbox{\hrule height #2 width \squaresize}}%
\rlap{\raise -.5\dd!
\vbox{\hrule height #3 width \squaresize}}%
\temp!=\squaresize\advance\temp! by .5\uu!
\rlap{\kern \squaresize \kern-.5\rr!
\vbox{\hrule height \temp! width#4 depth .5\dd!}}%
\rlap{\kern .5\squaresize\raise .5\squaresize
\vbox to 0pt{\vss\hbox to 0pt{\hss $#5$\hss}\vss}}%
\ht0=0pt \dp0=0pt \box0
\def\vsq!#1#2#3#4#5\endvsq!{\vbox
to \squaresize{\hrule width\squaresize height 0pt%
\vss\sq!{#1}{#2}{#3}{#4}{#5}}}
\newdimen \LL! \newdimen \UU! \newdimen \DD! \newdimen \RR!
\def\vvsq!{\futurelet\next\vvvsq!}
\def\vvvsq!{\relax
\ifx \next l\LL!=\Thickness \let\continue!=\skipnexttoken!
\else\ifx\next u\UU!=\Thickness \let\continue!=\skipnexttoken!
\else\ifx\next d\DD!=\Thickness \let\continue!=\skipnexttoken!
\else\ifx\next r\RR!=\Thickness \let\continue!=\skipnexttoken!
\else\ifx\next P\let\continue!=\place!
\else\def\continue!{\vsq!\LL!\UU!\DD!\RR!}%
\fi\fi\fi\fi\fi
\continue!}
\def\skipnexttoken!#1{\vvsq!}
\def\place! P#1#2#3{%
\rlap{\kern.5\squaresize\temp!=.5\squaresize\kern#1\temp!
\temp!=\squaresize
\advance\temp! by #2\squaresize \temp!=.5\temp!
\raise\temp!\vbox
to 0pt{\vss\hbox to 0pt{\hss$#3$\hss}\vss}}\vvsq!}
\def\Young#1{\LL!=\thickness \UU!=\thickness
\DD! = \thickness \RR! = \thickness
\vbox{\smallskip\offinterlineskip
\halign{&\vvsq! ## \endvsq!\cr #1}}}
\def\Youngt#1{\LL!=\thickness \UU!=
\thickness \DD! = \thickness \RR! = \thickness
\vtop{\offinterlineskip
\halign{&\vvsq! ## \endvsq!\cr #1}}}
\def\omit\hskip\squaresize{\omit\hskip\squaresize}
\catcode `!=12
\thickness=.4pt
We consider dimer coverings of an $m\times n$ rectangle,
with $m$ and $n$ even. We assign a vertical domino from row
$i$ to row $i+1$ the weight
$\sqrt {y_i}$ and a horizontal domino from column $j$ to
column $j+1$ the weight $\sqrt {x_j}$. For example, the
covering
$$\Thickness=0pt
\squaresize=30pt
\Young{dP0{-1}{\sqrt{y_1}}&rP10
{\sqrt{x_2}}&l&dP0{-1}
{\sqrt{y_1}}&
rP10{\sqrt{x_5}}&l&rP10{\sqrt{x_7}}
&l&dP0{-1}{\sqrt{y_1}}&dP0{-1}{\sqrt{y_1}}\cr
u&rP10{\sqrt{x_2}}&l&u&rP10
{\sqrt{x_5}}&l&rP10{\sqrt{x_7}}&l&u&u\cr}
$$
for $m=2$ and $n=10$ has weight $y_1^2
x_2x_5x_7$. (The weight will always be a product
of integral powers of the $x_i$ and $y_j$.)
Now I'll define what I call ``dimer tableaux.''
Take an $m/2$ by $n/2$ rectangle and split it
into two parts by a path from the lower left
corner to the upper right corner. For example
(with $m=6$ and $n=10$)
$$\Thickness=1pt
\squaresize 15pt
\Young{&&&&r\cr
&&&&lu\cr
d&dr&u&u&\cr}
$$
Then fill in the upper left
part with entries from 1, 2, \dots, $n-1$ so that
for adjacent entries
$\squaresize 10pt\lower 2pt\vbox{\Young{\mathstrut{\scriptstyle i}&\mathstrut{\scriptstyle j}\cr}}$
we have $i<j-1$ and for adjacent entries
$\squaresize 10pt\lower 8pt\vbox{\Young{\mathstrut{\scriptstyle i}\cr\mathstrut{\scriptstyle j}\cr}}$ we have $i\le j+1$, and fill in the
lower-right partition with entries from
$1,2\ldots, m-1$ with the reverse inequalities (
$\squaresize 10pt\lower 2pt\vbox{\Young{\mathstrut{\scriptstyle i}&\mathstrut{\scriptstyle j}\cr}}$
implies $i\le j+1$ and
$\squaresize 10pt\lower 8pt\vbox{\Young{\mathstrut{\scriptstyle i}\cr\mathstrut{\scriptstyle j}\cr}}$
implies $i<j-1$). We weight an $i$ in
the upper-left part by $x_i$ and a $j$ in the
lower-right part by $y_j$.
\begin{theorem}
The sum of the
weights of the $m\times n$ dimer coverings is
equal to the sum of the weights of the $m/2\times
n/2$ dimer tableaux.
\end{theorem}
My proof is not very enlightening; it essentially
involves showing that both of these are counted
by the same formula.
\begin{problem}
\label{prob20}
Is there an ``explanation'' for
this equality? In particular, is there a
reasonable bijective proof? Notes:
\begin{enumerate}
\item[(1)] The case $m=2$ is easy: the $2\times
10$ dimer covering above corresponds to the
$1\times 5$ dimer tableau
$$\squaresize
20pt\Thickness1pt\Young{dx_2&dx_5&drx_7&uy_1&uy_1\cr}$$
(there's only one possibility!).
\item[(2)] If we set $x_i=y_i=0$ when $i$ is even
(so that every two-by-two square of the dimer
covering may be chosen independently), then the
equality is equivalent to the identity
$$\prod_{i,j}(x_i+y_j)=\sum_{\lambda}s_{\lambda}(x)s_{\tilde
\lambda'}(y);$$
compare \cite[p.~37]{Macdonald}.
This
identity can be proved by a variant of
Schensted's correspondence, so a bijective proof
of the general equality would be essentially a
generalization of Schensted. Several people have
looked at the problem of a Schensted
generalization corresponding to the case in which
$y_i=0$ when
$i$ is even.
\item[(3)] The analogous results in which $m$ or
$n$ is odd are included in the case in which $m$
and $n$ are both even. For example, if we take
$m=4$ and set $y_3=0$, then the fourth row of a
dimer covering must consist of $n/2$ horizontal
dominoes, which contribute
$\sqrt{x_1x_3\cdots x_{n-1}}$ to the weight, so
we are essentially looking at dimer coverings
with three rows.
\end{enumerate}
\end{problem}
\begin{progress}
A special case of the Robinson--Schensted algorithm given by
Sundquist et al.\ \citeyear{PrSWW}
can be used to get a bijection for a special case of the problem, in
which one sets $y_i = 0$ for all $i$ even, so that we are looking at
dimer coverings (or domino tilings) in which every vertical domino goes
from row $2i+1$ to row $2i+2$ for some $i$. These tilings are not very
interesting because they break up into tilings of 2-by-$n$ rectangles.
But even so, the Robinson--Schensted bijection is nontrivial.
\end{progress}
\section{New Problems}
Let $N(a,b)$ denote the number of matchings
of the $a$-by-$b$ rectangular grid.
Kasteleyn showed that $N(a,b)$ is equal to the square root of
the absolute value of
$$\prod_{j=1}^{a} \prod_{k=1}^{b}
\left(2 \cos \frac{\pi j}{a+1} + 2i \cos \frac{\pi k}{b+1}\right).$$
Some number-theoretic properties of $N(a,b)$ follow from this representation
(see, e.g., \cite{PrCo})
but lack a combinatorial explanation.
The next two problems describe two such facts.
\begin{problem}
\label{prob21}
Give a combinatorial proof of the fact that
$N(a,b)$ divides $N(A,B)$ whenever
$a+1$ divides $A+1$ and $b+1$ divides $B+1$.
\end{problem}
\begin{progress}
Bruce Sagan has given an answer in the ``Fibonacci case'' $a=2$.
A matching of a $2$-by-$(kn-1)$ grid
either splits up as a matching of a $2$-by-$(n-1)$ grid on the left
and a $2$-by-$(kn-n)$ grid on the right
or it splits up as a matching of a $2$-by-$(n-2)$ grid on the left,
a horizontal matching of a $2$-by-$2$ grid in the middle,
and a matching of a $2$-by-$(kn-n-1)$ grid on the right.
Hence
$$N(2,@kn-1) = N(2,@n-1)@N(2,@kn-n) + N(2,@n-2)@N(2,@(k-1)n-1).$$
{}From this formula one can prove that $N(2,n-1)$ divides $N(2,kn-1)$
by induction on $k$.
Volker Strehl has approached the problem in a different way;
his ideas make it seem likely
that a better combinatorial understanding of resultants,
in combination with known interpretations of Chebyshev polynomials,
would be helpful in approaching this problem.
\end{progress}
\begin{problem}
\label{prob22}
Give a combinatorial proof of the fact that
$N(a,2a)$ is always congruent to 1 mod 4.
\end{problem}
\hskip\parindent
(Pachter \citeyear{PrPa} has demonstrated the sort of combinatorial
methods one can use in such problems.)
\medbreak
Even without Kasteleyn's formula,
it is easy to show (e.g., via the transfer-matrix method)
that for any fixed $a$,
the sequence of numbers $N(a,b)$ (with $b$ varying)
satisfies a linear recurrence relation with constant coefficients.
Indeed, consider all $2^a$ different ways
of removing some subset of the $a$ rightmost vertices in the $a$-by-$b$ grid;
this gives us $2^a$ ``mutilated'' versions of the graph.
We can set up recurrences that link
matchings of mutilated graphs of width $b$
with matchings of mutilated graphs of width $b$ and $b-1$,
and standard algebraic methods allow us to turn this system
of joint mutual recurrences of low degree
into a single recurrence of high degree
governing the particular sequence of interest,
which enumerates matchings of unmutilated rectangles.
The recurrence obtained in this way is not, however, best possible,
as one can see even in the simple case $a=2$.
\begin{problem}[Stanley]
Prove or disprove that the minimum degree of a linear recurrence
governing the sequence $N(a,1),N(a,2),N(a,3),\dots$
is $2^{\lfloor (a+1)/2 \rfloor}$.
\end{problem}
\begin{progress}
Observations made by Stanley \citeyear[p.~87]{PrSt2} imply that
the conjecture is true when $a+1$ is an odd prime.
\end{progress}
\hskip\parindent
The idea of mutilating a graph by removing some vertices along its boundary
leads us to the next problem. It has been observed for small values of $n$
that if one removes equal numbers of black and white vertices from the
boundary of a $2n$-by-$2n$ square grid, the number of matchings
of the mutilated graph is less than the number of matchings of
the original graph. In fact, it appears to be true that
one can delete \textit{any} subset of the vertices of the square grid and
obtain an induced graph with strictly fewer matchings than the original.
It is worth pointing out that not every graph shares this property with
the square grid. For instance,
if $G$ is the Aztec diamond graph of order 5 and $G'$ is the graph obtained
from $G$ by deleting the middle vertices along the northwest and northeast
borders, then $G$ has 32768 matchings while $G'$ has 59493.
\begin{problem}
Prove or disprove that every subgraph of the $2n$-by-$2n$ grid graph
has strictly fewer matchings.
\end{problem}
\hskip\parindent
Next we come to a variant on the Aztec dungeon region shown in Figure 4.
Figure 10 shows an ``hexagonal dungeon'' with sides $2,4,4,2,4,4$.
Matt Blum's investigation of these shapes has led him to discover
many patterns; the most striking of these patterns forms the basis
of the next problem.
\begin{figure}
\centerline{\psfig{file=\figdir/hexdungeon,scale=78}}
\caption{An hexagonal dungeon.}
\label{hexdungeon}
\end{figure}
\begin{problem}
Show that the hexagonal dungeon with sides $a,2a,b,a,2a,b$ has exactly
$$13^{2a^2} 14^{\lfloor a^2/2 \rfloor}$$
diform tilings, for all $b \geq 2a$.
\end{problem}
\hskip\parindent
Unmatchable bipartite graphs can sometimes give rise to
interesting quasimatching problems, either by way of $K K^*$
(see Problem 6) or by systematic addition or deletion of vertices or edges.
The former sort of problem simply asks for the determinant
of $K K^*$ (where we may assume that $K$ has more columns than rows).
When the underlying graph has equal numbers of black and white
vertices, this is just the square of the number of matchings,
but when $K$ is a rectangular matrix, $K K^*$ will in general
have a nonzero determinant, even though the graph has no
matchings.
\setcounter{topnumber}{1}
\begin{problem}
Calculate the determinant of $K K^*$ where $K$ is the Kasteleyn--Percus
matrix of the $a,b,c,d,e,f$ honeycomb graph.
\end{problem}
\hskip\parindent
(Note that in this case we can simply take $K$ to be the bipartite
adjacency matrix of the graph.)
Cases of special interest are $a$, $b+1 $, $c $, $a+1 $, $b $, $c+1$
and $a,b,a,b,a,b$ hexagons. These two cases overlap in the
one-parameter family of $a $, $a+1 $, $a $, $a+1 $, $a $, $a+1$
hexagons. For instance, in the case of the $3,4,3,4,3,4$ hexagon,
$\det(K K^*)$ is $2^8\cdot 3^3\cdot 7^6$.
\begin{problem}
\label{prob27}
Calculate the determinant of $K K^*$ where $K$ is the Kasteleyn--Percus
matrix of an $m$-by-$n$ Aztec rectangle, or where $K$ is the
Kasteleyn--Percus matrix of the ``fool's diamond'' of order $n$.
(The fool's diamond of order 3 is the following region:
$$\vcenter\brd{
X
XXX
XXXXX
XXX
X
}$$
Fool's diamonds of higher orders are defined in a similar way.)
\end{problem}
\begin{progress}
In the case of Aztec rectangles,
Matt Blum has found general formulas for $\det(KK^*)$ when
$m$ is 1, 2, or 3. For fool's diamonds, we get
$$
\medmuskip 1mu
\displaylines{
1\cr 2\cr 3\cdot 5\cr 2^7\cdot 3\cr 3^2\cdot 5^3\cdot 29\cr
2^9\cdot 3\cdot 5 \cdot 7 \cdot 13^2\cr
7^3\cdot 13^4\cdot 29^2\cr 2^{25}\cdot 3\cdot 7^2\cdot 17^3\cr}
$$
(One might also look at ``fool's rectangles''.)
\end{progress}
\hskip\parindent
Another thing one can do with an unmatchable graph is add extra edges.
Even when this ruins the bipartiteness of the graph, there can still
be interesting combinatorics. For instance, consider the $2,4,2,4,2,4$
hexagon-graph; it has an even number of vertices, but it has a surplus
of black vertices over white vertices. We therefore introduce
edges between every black vertex and the six nearest black vertices.
(That is, in each hexagon of the honeycomb, we draw a triangle
connecting the three black vertices, as in Figure~\ref{hextwo}.)
Then the graph has $5187 = 3\cdot7\cdot13\cdot19$ matchings.
\begin{figure}
\centerline{\psfig{file=\figdir/hextwo,scale=90}}
\caption{A hexagon with extra edges.}
\label{hextwo}
\end{figure}
\begin{problem}
\label{prob28}
Count the matchings of the $a,b,c,d,e,f$
hexagon-graph in which extra edges have been drawn connecting
vertices of the majority color.
\end{problem}
\begin{figure}
\centerline{\psfig{file=\figdir/rectwo,scale=96}}
\caption{An Aztec rectangle with extra edges.}
\label{rectwo}
\end{figure}
\begin{figure}
\centerline{\psfig{file=\figdir/holtwo,scale=75}}
\caption{A holey Aztec rectangle with extra edges.}
\label{holtwo}
\end{figure}
\hskip\parindent
What works for honeycomb graphs works (or seems to work)
for square-grid graphs as well. If one adds edges joining each vertex
of majority color to the four nearest like-colored vertices
in the $n$ by $n+2$ Aztec rectangle graph as in Figure~\ref{rectwo},
one gets a graph for which the number of matchings grows like
$2^2\cdot3$, $2^3\cdot3\cdot7$, $2^7\cdot 3\cdot 11$, $2^{17}\cdot 5\cdot 31$,
etc. If one does the same for the holey $2n-1$ by $2n$ Aztec rectangle
from which the central vertex has been removed, as in Figure~\ref{holtwo},
one gets the numbers
$2^6\cdot 7$, $2^9\cdot 3^2\cdot 13\cdot 17$, $2^{23}\cdot 5^3\cdot 31$, etc.
\begin{problem}
\label{prob29}
Count the matchings of the $a$ by $b$ Aztec
rectangle (with $a+b$ even) in which extra edges have been drawn
connecting vertices of the majority color. Do the same for the
$2n-1$ by $2n$ holey Aztec rectangle.
\end{problem}
\hskip\parindent
Other examples of nonbipartite graphs for which the number of matchings
has only small prime factors arise when one takes the quotient of a
symmetrical bipartite graph modulo a symmetry-group at least one element
of which interchanges the two colors; Kuperberg \citeyear{PrKu1} gives
some examples of this.
In general, there seem to be fewer product-formula enumerations of
matchings for nonbipartite graphs than for bipartite graphs.
Nevertheless, even in cases where no product formula has been found,
there can be patterns in need of explanation.
Consider the one-parameter family of graphs illustrated in Figure 14
for the case $n=7$ (based on the same nonbipartite infinite graph
\begin{figure}
\centerline{\psfig{file=\figdir/diab,width=2.5in}}
\caption{An isosceles right triangle graph with extra edges.}
\label{diab}
\end{figure}
as Figures 12 and 13). Such a graph has an even number of vertices
whenever $n$ is congruent to 0 or 3 modulo 4.
Here are the data for the first few cases, courtesy of Matt Blum:
$$
\medmuskip 1mu
\tabskip0pt plus 1fil
\vbox{
\halign to \hsize{\hfil$#$&\hfil#&$#$\hfil\cr
n& number of matchings& $factorization$\cr
3& 3& 3\cr
4& 6& 2 \cdot 3\cr
7& 1065& 3 \cdot 5 \cdot 71\cr
8& 6276& 2^2 \cdot 3 \cdot 523\cr
11& 45949563& 3^2 \cdot 11 \cdot 464137\cr
12& 807343128& 2^3 \cdot 3^2 \cdot 1109 \cdot 10111\cr
15& 221797080594801& 3^2 \cdot 24644120066089\cr
16& 11812299253803024& 2^4 \cdot 3 \cdot 246089567787563\cr
19& 117066491250943949567763& 3 \cdot 89 \cdot 28289
\cdot 15499002371714201\cr
20& 19100803250397148607852640& 2^5 \cdot 3^2 \cdot 5 \cdot 41 \cdot 367
\cdot 881534305952328473\cr
}}$$
The following problem describes some of Blum's conjectures:
\begin{problem}
\label{prob30}
Show that for the isosceles right triangle graph
with extra edges, the number of matchings is always a multiple of 3.
Furthermore, show that the exact power of 2 dividing the number of
matchings is $2^{n/4}$ when $n$ is 0 modulo 4, and $2^0(=1)$ when
$n$ is 3 modulo 4.
\end{problem}
\hskip\parindent
This property of divisibility by 3 pops up in another problem of
a similar flavor. Consider the graph shown in Figure 15, which is
just like the one shown in Figure 9, except that half of the triangular
cells have an extra vertex in them, connected to the three nearest
vertices. (Note also the resemblance to Figure~11.)
\begin{figure}
\centerline{\psfig{file=\figdir/newtri,width=1.4in}}
\caption{An equilateral triangle graph with extra vertices and edges.}
\label{newtri}
\end{figure}
\begin{problem}
\label{prob31}
Show that for the equilateral triangle graph
with extra vertices and edges, the number of matchings is always a
multiple of 3.
\end{problem}
\hskip\parindent
(I refrain from making a conjecture about the exponent of 2, though
the data contain patterns suggestive of a general rule.)
\medskip
It may be too soon to try to assemble into one coherent picture all
the diverse phenomena discussed in the preceding 31 problems. But
I have noticed a gratuitous symmetry that governs many of the exact
formulas, and I will close by pointing it out. Consider, for example,
the MacMahon--Macdonald product
$$M_n = \prod_{i=0}^{n-1} \prod_{j=0}^{n-1} \prod_{k=0}^{n-1}
\frac{i+j+k+2}{i+j+k+1}$$
that counts matchings of the $n,n,n$ semiregular honeycomb graph.
We find that the ``second quotient'' $M_{n-1} M_{n+1} / M_n^2$
is the rational function
$$\frac{27}{64} \,\frac{(3n-2)(3n-1)^2(3n+1)^2(3n+2)}{(2n-1)^3(2n+1)^3}$$
which is an even function of $n$.
The right hand side in Bo-Yin Yang's theorem (giving the number of diabolo
tilings of a fortress of order $n$) has a power of 5 whose exponent is
$n^2/4$ when $n$ is even and $(n^2-1)/4$ when $n$ is odd; this too is
an even function of $n$.
Domino tilings of Aztec diamonds are enumerated by the formula
$2^{n(n+1)}$. Here the symmetry is a bit different:
replacing $n$ by $-1-n$ leaves the answer unaffected.
The right hand side of Mihai Ciucu's theorem (giving the number of
diform tilings of an Aztec dungeon of order $n$) has a power of 13
whose exponent is
$(n+1)^2/3$ or $n(n+2)/3$ (according to whether or not $n$ is 2 mod 3).
so that the symmetry corresponds to replacing $n$ by $-2-n$.
There are other instances of this kind that arise,
in which some base is raised to the power of
some quadratic function of $n$;
in each case, the quadratic function
admits a symmetry that preserves the integrality of $n$
(unlike, say, the quadratic function $n(3n+1)/2$,
which as a function from integers to integers
does not possess such a symmetry).
\begin{problem}
\label{prob32}
For many of our formulas,
the ``algebraic'' (right hand) side
is invariant under substitutions that make the
``combinatorial'' (left hand) side
meaningless, insofar as one cannot speak
of graphs with negative numbers of vertices or edges.
Might this invariance nonetheless have some deeper
significance?
\end{problem}
\hskip\parindent
Cohn \citeyear{PrCo} has found
another example of gratuitous symmetry related to tilings.
\section*{Acknowledgements}
This research was conducted with the support of
the National Science Foundation, the National Security Agency,
and the M.I.T. Class of 1922 Career Development fund.
I am deeply indebted to the past and present members
of the Tilings Research Group
for their many forms of assistance:
Pramod Achar, Karen Acquista, Josie Ammer, Federico Ardila,
Rob Blau, Matt Blum, Carl Bosley, Ruth Britto-Pacumio, Constantin Chiscanu,
Henry Cohn, Chris Douglas, Edward Early, Nicholas Eriksson, David
Farris, Lukasz Fidkowski, Marisa Gioioso,
David Gupta, Harald Helfgott, Sharon Hollander, Dan Ionescu,
Sameera Iyengar, Julia Khodor, Neelakantan Krishnaswami,
Eric Kuo, Yvonne Lai, Ching Law, Andrew Menard, Alyce Moy,
Anne-Marie Oreskovich, Ben Raphael,
Vis Taraz, Jordan Weitz, Ben Wieland, Lauren Williams,
David Wilson, Jessica Wong, Jason Woolever, and Laurence Yogman.
I also acknowledge the helpful comments on this manuscript
given by Henry Cohn and Richard Stanley,
and the information provided by Jerry Dias, Michael Fisher and Horst Sachs
concerning the connections between matching theory and the physical sciences.
\bibitemsep=5pt plus 1pt
\def\citeauthoryear #1#2#3{#2\ #3}
\bibliographystyle{msribib}
|
\section{Introduction}
In recent years, there has been extensive experimental and theoretical
work in the field of mesoscopic systems, including superconducting
structures in proximity with normal metals \cite{review}. In
particular, the paramagnetic reentrance phenomenon \cite{visaniprl}
has received a wide interest, mainly because of being promoted by the
recent understanding of the high-temperature diamagnetic response of
rather clean normal-metal--superconductor (NS) proximity structures
\cite{bmueller1} in the context of the quasiclassical Eilenberger
theory including elastic scattering \cite{belzig}. In this Letter, we
discuss the very low temperature reentrant behavior of two of the AgNb
samples of Ref.\ \cite{bmueller1}, covering a larger mesoscopic regime
with respect to previous measurements \cite{visaniprl}.
Recently, two Letters \cite{bruder,fauchere99} have addressed the
origin of paramagnetic currents in NS systems, which might lead to an
understanding of the paramagnetic reentrance phenomenon. The work of
Bruder and Imry \cite{bruder} is based on the presence of
non-Andreev-reflecting semiclassical trajectories at the outer surface
of a nonsingly connected proximity system (glancing states), which
carry predominantly paramagnetic currents. This work has been subject
to debate because of the small magnitude \cite{comment}. A different,
more elaborate approach by Fauch\`ere \textit{et al}.
\cite{fauchere99} assumes a net repulsive interaction in the noble
metals. The $\pi$ shift of the order parameter at the NS interface
then leads to a paramagnetic instability of Andreev pairs.
The first work reflects the cylindrical geometry of our NS system, but
it does not address the experimental signatures of the paramagnetic
reentrance, namely absolute value of order 1, temperature dependence,
nonlinearity, hysteresis, or dissipation \cite{visaniprl}. The latter
three features might give evidence for a spontaneous magnetization in
the samples, as proposed in Ref.\ \cite{fauchere99}. However, the
second theoretical approach \cite{fauchere99} does not reach beyond
qualitative accordance with our experiment.
Here we discuss an investigation of the paramagnetic reentrant effect
extending to the $\mu$K region. By covering five decades in
temperature, we have been able to extract the correct temperature
dependence of the NS proximity structure below $T_{c}$. Over a large
mesoscopic regime, the cylindrical structure clearly displays
different levels of coherence along integer multiples of the wire
perimeter $L$. Unfortunately, neither of the two present theories
\cite{bruder,fauchere99}, which discuss different sources of
paramagnetism in NS structures, obtain the correct absolute value as
well as the $T$ dependence of the paramagnetic reentrant effect.
The two samples reported here are ensembles of cylindrical wires with a
superconducting core of soft niobium ($RRR\approx 300$) concentrically
embedded in a normal-metal matrix of 6N silver.
Their total diameter was mechanically reduced by several steps of
swagging and co-drawing \cite{flukiger} to final values
$41\,\mu{\mathrm m}$ and $23\,\mu{\mathrm m}$, with normal layer thicknesses
$d_{N}=5.5\,\mu{\mathrm m}$ and $3.3\,\mu{\mathrm m}$ (the ratio
$d_{N}/L$ is approximately the same), respectively.
The samples were
\begin{figure}
\includegraphics[width=0.96\linewidth]{XdclogT.eps}
\caption{Magnetic susceptibility $\chi_{N}(T)$, between
$150\,\mu{\mathrm K}$ and $9\,{\mathrm K}$. For both samples we
show $\chi_{ac}(T)$ ($+$) and $\chi_{dc}(T)$ ($\bullet$). The
arrows mark the direction of $T$--changes. }
\protect\label{Xdc(logT)}
\end{figure}
\noindent annealed after the last drawing,
and values of the mean free path $\ell_{N}\sim(0.5$--$0.8)d_{N}$
were obtained. For more details see
Refs.\ \cite{journlowtempphys,bmueller1}.
Extensions of the measurements to $\mu$K temperatures were performed
at the ultralow temperature (ULT) facility at the University of
Bayreuth. There, an experimental setup was installed for inductive
measurements using an rf--SQUID sensor. Magnetic fields were applied
along the axis of the wires.
For the ULT experiments, we took parts of the wire bundle measured in
our dilution refrigerator \cite{visaniprl,bmueller1}, and glued them
with GE 7031 varnish to high purity gold foils tightly attached to a
silver finger, in good electrical contact with the Cu demagnetization
stage \cite{gloos88}. Thus, about 200 wires were mounted.
Temperatures were measured with a Pt pulsed NMR thermometer
\cite{gloos88}.
In the following we report on the temperature dependent magnetic
susceptibility of our relatively clean silver-niobium samples 3AgNb
[$d_{N}=5.5\,\mu{\mathrm m}$] and 5AgNb [$d_{N}=3.3\,\mu{\mathrm m}$].
In Fig.\ \ref{Xdc(logT)} the total magnetic susceptibility is shown as
a function of temperature. We show $\chi_{ac}(T)$ between
$4\,{\mathrm mK}$ and $9\,{\mathrm K}$, measured in our dilution
refrigerator with field amplitude $H_{ac}=33\,{\mathrm mOe}$ and
frequency $\nu=80\,{\mathrm Hz}$, as well as $\chi_{dc}(T)$ at
constant $H_{dc}$ at ULT and LT.
At temperatures below the critical temperature of Nb
$T_{c}=9.2\,{\mathrm K}$, the magnetic susceptibility of the N layer
exhibits diamagnetism induced through Andreev reflection at the highly
transparent NS interface. At lower temperatures, it develops almost
total Meissner screening in the Ag layer \cite{bmueller1}. Below
$T_{r}\sim 100\,{\mathrm mK}$ the signature of reentrance is observed
both in $\chi_{ac}(T)$ \cite{visaniprl} and $\chi_{dc}(T)$, with the
development of an additional paramagnetic susceptibility
$\chi_{\mathrm para}(T)$, such that $\chi_{N}(T)=\chi_{\mathrm dia}(T)+\chi_{\mathrm para}(T)$.
For sample 3AgNb, the susceptibility $\chi_{N}$ saturates
below $T^{\mathrm sat}\approx 400\,\mu{\mathrm K}$, displaying a
complete cancellation of the induced diamagnetic susceptibility in N,
such that only the diamagnetism in S seems to remain.
For sample 5AgNb the susceptibility $\chi_{N}$ shows
saturation below $T^{\mathrm sat}\approx 800\,\mu{\mathrm K}$
at a paramagnetic value $4\pi\chi_{N}^{\mathrm
sat}\approx 1$, indicating a complete cancellation of the
\textit{total} diamagnetic susceptibility in N plus S.
The anomalously large magnitude of the paramagnetic reentrant
susceptibility $\chi_{\mathrm para}$ at ultralow temperatures is rather
intriguing. This magnitude, and particularly its dependence on the
sample size, is not explained by existing theories. At this moment,
it is not clear, if paramagnetic reentrance is (i) an intrinsic effect
of mesoscopic NS proximity structures in the very low temperature
limit or (ii) the result of two independent phenomena. In the latter
case the selection of the NS materials could be important.
A detailed inspection $\chi_{N}$ around its minimum reveals that, the
reentrance temperature $T_{r}$ is decreased under a field
$H_{dc}\approx 0.2\,{\mathrm Oe}$. The ac and dc curves show reentrant
temperatures $T_{r}=97\,{\mathrm mK}$ and $T_{r}^{H}=83\,{\mathrm mK}$
for sam-
\begin{figure}
\includegraphics[width=0.96\linewidth]{XdcTlog.eps}
\caption{Reentrant paramagnetic susceptibility $\chi_{\mathrm
para}(T)$ below $T_{r}$. For both samples two measurements of
$\chi_{dc}(T)$ ($\bullet$) are shown, and $\chi_{ac}(T)$
($+$). The thick arrows indicate the direction of temperature
changes. The temperature scales are $\propto 1/L$. The
vertical arrows indicate the temperatures at which the
condition $\xi_{N}(T)=nL$ is met. }
\protect\label{Xdc(T)log}
\end{figure}
\noindent ple 3AgNb ($T_{r}=149\,{\mathrm mK}$ and
$T_{r}^{H}=113\,{\mathrm mK}$ for sample 5AgNb), respectively.
Neglecting the weak-field effect on $T_{r}$, $\chi_{dc}(T)$ matches
$\chi_{ac}(T)$ for sample 3AgNb (5AgNb) between $15\,{\mathrm mK}$
($30\,{\mathrm mK}$) and $1\,{\mathrm K}$. This is rather noticeable,
considering the strong nonlinearity in magnetic field discussed below.
Above $T\approx1\,{\mathrm
K}$, $\chi_{dc}(T)$ deviates from $\chi_{ac}(T,H_{ac}\approx 0)$, due
to the depression of the weak induced Andreev pair potential by finite
fields. Below $15\,{\mathrm mK}$ ($30\,{\mathrm mK}$), $\chi_{dc}(T)$
agrees with $\chi_{ac}(T)$ only on measurements done under cooling.
Below $15\,{\mathrm mK}$ ($30\,{\mathrm mK}$) the warming up curve of
$\chi_{dc}(T)$ lies above the cooling curves, corresponding to stronger
paramagnetism adding to the full Meissner screening.
In our arrangement, samples are in rather good thermal contact with
the Cu demagnetization stage and the Pt NMR thermometer. Indeed, the
measured thermal relaxation times at the lowest temperatures remain
below $1000\,{\mathrm s}$. Nevertheless, the susceptibility shows
hysteresis. We have found the highest levels of paramagnetic
reentrance only after allowing the NS system to remain at much below
their saturation temperatures for long periods of time (one week or
longer).
Fig.\ \ref{Xdc(T)log} shows the reentrant paramagnetic susceptibility
$\chi_{\mathrm para}(T)$ below $T_{r}$. The data exponentially
increases as $\chi_{\mathrm para}(T)=A\exp{(-T/T^\ast)}$. For sample
3AgNb [Fig.\ \ref{Xdc(T)log}(a)], the prefactor $A_{1}=0.24$ and
characteristic temperature $T_{1}^\ast=13\,{\mathrm mK}$ were obtained
from $\chi_{ac}(T)$. The cooling and warming $\chi_{dc}(T)$, above
$\approx 13\,{\mathrm mK}$ reproduce well the behavior of
$\chi_{\mathrm para}$ observed in $\chi_{ac}(T)$. Near $\approx
13\,{\mathrm mK}$, $\chi_{dc}$ displays a kink, leading to an
approximate doubling of the logarithmic slope, with
$T^\ast_{2}=5.5\,{\mathrm mK}\sim T^\ast_{1}/2$ and a prefactor
$A_{2}=1.1$.
For sample 5AgNb [Fig.\ \ref{Xdc(T)log}(b)], the prefactor $A_{1}=0.4$
and characteristic temperature $T_{1}^\ast=20\,{\mathrm mK}$ were
obtained from $\chi_{ac}(T)$. Again the cooling and warming
$\chi_{dc}(T)$ above $\approx 30\,{\mathrm mK}$ follow the
$T$ behavior of $\chi_{ac}(T)$. The curves show a kink in the
susceptibility, displaying a doubling of the logarithmic slope, with
$T^\ast_{2}=11\,{\mathrm mK}\sim T^\ast_{1}/2$ for the second line.
In addition, around $\approx14\,{\mathrm mK}$, a second doubling of
the logarithmic slope of $\chi_{dc}(T)$ occurs, with
$T^\ast_{3}=4.8\,{\mathrm mK}\sim T^\ast_{2}/2$ for the third line.
The coherence length of the Andreev pairs $\xi_{N}$ in Ag, obtained
from our breakdown field measurements, is in agreement with the clean
limit theory, $\xi_{N}=\hbar v_{F}/2\pi k_{B}T=1.69\,\mu{\mathrm
m}/T({\mathrm K})$ \cite{bmueller1}. At the temperature of the first
kink in $\chi_{\mathrm para}$, $\xi_{N}(T)$ reaches approximately the
value of a single wire's circumference $L=130\,\mu{\mathrm m}$
($L=72\,\mu{\mathrm m}$). In Fig.\ \ref{Xdc(T)log} we have indicated
by vertical arrows the temperature at which the equality
$\xi_{N}(T)=L$ is met. For sample 5AgNb, at the temperature of the
second kink it is $\xi_{N}(T)=2L$. The values of $T_{1}^\ast$,
$T_{2}^\ast$, and $T_{3}^\ast$, as well as the position of the kinks,
which are located approximately at $T_{1}^\ast$, $T_{2}^\ast$, give
evidence for different levels of quantum coherence on the mesoscopic
length scale $L$. The temperatures $T^\ast$, which characterize the
different levels as well as the kinks can be written as
$T^\ast\approx\hbar v_{F}/2\pi k_{B}nL$, or in the equivalent form
$\xi_{N}(T^\ast)=nL$, with $n=1,2,4$. The reentrant paramagnetic
susceptibility then is
\begin{equation}
\label{eq1}
\chi_{\mathrm para}=A_{n} \exp{\left[-\frac{nL}{\xi_{N}(T)}\right]},\quad {\mathrm with}\, n=1,2,4\, .
\end{equation}
This characteristic behavior is not obtained by present theories. The
correct theory of the paramagnetic reentrant effect should describe
the suscptibility in accordance with Eq.\ \ref{eq1}.
For a theoretical understanding of the origin of the reentrant effect,
it is important to investigate the susceptibility under fields, as
well as the magnetization. In the following we discuss isothermal ac
susceptibility measurements of sample 5AgNb, as a function of magnetic
field [Figs.\ \ref{X(H,[T])nh} and \ref{XacMdc(H)nh}(a)]. The
isothermal susceptibility $\chi_{N}(H)$ shows nonlinear behavior in
the entire field regime. At $7\,{\mathrm mK}$ two curves are shown,
the first measurement directly after cooldown, and the second one
\begin{figure}
\includegraphics[width=0.96\linewidth]{XHT.eps}
\caption{Isothermal ac-magnetic susceptibility $\chi_{N}(H)$,
below $50\,{\mathrm mK}$.
The arrows indicate the direction of field changes.}
\protect\label{X(H,[T])nh}
\end{figure}
\begin{figure}
\includegraphics[width=0.96\linewidth]{XacMdcH.eps}
\caption{(a) Nonlinear ac susceptibility $\chi_{N}(H)$
and (b) isothermal dc magnetization $M_{dc}(H)$ in full field
cycles, starting with positive values;
the black lines guide the eye.}
\protect\label{XacMdc(H)nh}
\end{figure}
\noindent $\sim 10^6\,{\mathrm s}$ after the first one. The lower of the two
curves at $7\,{\mathrm mK}$ shows pronounced hysteresis of the
reentrant part at fields below $20\,{\mathrm Oe}$. Near 20\,Oe the
specimen screens the magnetic flux most effectively, before the well
known magnetic breakdown transition occurs
\cite{bmueller1,journlowtempphys,belzig:96,fauchere}. For this
sample, the transition is nearly temperature independent for
$T<50\,{\mathrm mK}$. However, the minimum susceptibility before the
breakdown is far from complete screening. At $T=7\,{\mathrm mK}$, it
reaches only 30\,$\%$ of $-1/4\pi$, due to the paramagnetic
contribution.
For very small dc magnetic fields, the magnetic susceptibility grows
rather steeply. At higher fields, that increase slows down, the
susceptibility reaching its maximum value at about $2.5\,{\mathrm
Oe}$, before turning towards less paramagnetic values. As the
temperature is increased, the curves show a less pronounced maximum at
about $2.5\,{\mathrm Oe}$ and reduced hysteresis. The detailed
behavior at very low fields differs from the results of similar
measurements, performed on a bigger wire bundle, reported in Ref.\
\cite{visaniprl}. Possibly, the different geometrical arrangement of
the wires in the bundles affect the average susceptibility. Indeed,
measurements on a single wire are desirable.
Cycling the magnetic field in both directions at $7\,{\mathrm mK}$
[Fig.\ \ref{XacMdc(H)nh}(a)], we observe two low-field peaks of
$\chi_{N}(H)$, which are displaced symmetrically from zero (residual
field $<2\,{\mathrm mOe}$). However, the curves are not symmetric
below $20\,{\mathrm Oe}$, displaying a reduction of hysteresis after
each half-cycle. Furthermore, after cooldown from above $50\,{\mathrm
mK}$, cycling the field at $7\,{\mathrm mK}$, and waiting for $\sim
10^6\,\mathrm{s}$, the low-field peak at about $2.5\,{\mathrm Oe}$
grows up. This can be observed in the upper curves in Figs.\
\ref{X(H,[T])nh} and \ref{XacMdc(H)nh}(a). After cycling the field
and waiting, the whole system crosses to a more stable state, with
more pronounced paramagnetic susceptibility.
Hysteresis and nonlinearity are also observed in dc magnetization
curves, e.g. shown for a full cycle at $7\,{\mathrm mK}$ in Fig.\
\ref{XacMdc(H)nh}(b). For the second half-cycle, nonlinearity becomes
less pronounced, in accordance with our findings in the susceptibility
curves. The measured magnetization lies between two lines at fields
below the breakdown transition. At low fields one observes clearly
the deviation of the magnetization from the induced Meissner
screening. At higher fields the curve asymptotically approaches the
drawn line indicating linear Meissner-like behavior plus a constant
\textit{paramagnetic} magnetization. The $H=0$ intersections of the
lines suggest a field-independent magnetization $4\pi M_{0}\approx
1\,{\mathrm G}$. It is interesting to notice that a spontaneous
magnetization ot the same order was found in the theoretical model of
Fauch\'ere \textit{et al}. under the assumption of a negative average
order parameter in N, $\Delta_{N}/k_{B}\approx-160\,{\mathrm mK}$
\cite{fauchere99}. Moreover, some features of the paramagnetic
reentrant effect, such as nonlinearity, hysteresis, and dissipation,
are in qualitative agreemnet with their results \cite{fauchere99}.
At this point, a direct comparison of our experiment with theory
\cite{bruder,fauchere99} is not possible. The magnitude of
$\chi_{\mathrm para}$ of order 1, the characteristic dependence of the
magnetic susceptibility on $T$ and $L$, as described by Eq.\
\ref{eq1}, as well as its field dependence remain to be obtained by
theory. More theoretical work and also more experiments are
necessary.
In summary, the paramagnetic reentrance phenomenon in AgNb cylinders
of high purity is a nonlinear effect of anomalously strong magnitude.
It shows strong deviations from induced Meissner screening in the
low-temperature--low-field corner of the $H$-$T$ phase diagram.
It displays dissipation \cite{visaniprl}, hysteresis, and long time
relaxation behavior.
In the mesoscopic regime, the exponential temperature dependence of
the magnetic susceptibility with characteristic temperature
$T^\ast\approx\hbar v_{F}/2\pi k_{B}nL$ has the fingerprint of several
levels of quantum coherence along integer multiples $n=1,2,4$ of the
wire perimeter $L$.
Paramagnetic reentrance has also been observed in other NS materials
\cite{visaniprl}. In our most recent experiments we have found
reentrant behavior in gold-niobium cylinders. This has to be viewed
in the light of expected superconductivity in Au below $T_{c}\approx
200\,\mu{\mathrm K}$ \cite{hoyt}. In consideration of this, the
origin of this puzzling paramagnetism in mesoscopic NS cylinders is
still an open question.
We wish to express our gratitude to R. Frassanito, M. Nider{\"o}st,
and P. Visani for their invaluable contributions to earlier
experimental work. We thank the group at the ULT facility in Bayreuth
for their help and support. We acknowledge discussions with W.
Belzig, G. Blatter, C. Bruder, A. Fauch\`ere, and Y. Imry. We
acknowledge partial support from the ``Schweizerischer Nationalfonds
zur F{\"o}rderung der Wissenschaftlichen Forschung'' and the
``Bundesamt f{\"u}r Bildung und Wissenschaft'' (EU Program ``Training
and Mobility of Researchers'').
|
\section{Introduction}
\newfont{\graf}{eufm10}
\newcommand{\mbox{\graf h}}{\mbox{\graf h}}
Let $X^3 (c)$ be a 3-dimensional space form of constant curvature $c=0$ or 1 and admitting a real Killing spinor with respect to some spin structure. Consider a compact, oriented and immersed surface $M^2 \subset X^3(c)$ with mean curvature $H$. The spin structure of $X^3 (c)$ induces a spin structure on $M^2$. Denote by $D$ the corresponding Dirac operator acting on spinor fields defined over the surface $M^2$. The first eigenvalue $\lambda_1^2 (D)$ of the operator $D^2$ and the first eigenvalue $\mu_1$ of the Schr\"odinger operator $\Delta + H^2 +c$ are related by the inequality
\[ \lambda_1^2 (D) \le \mu_1 (\Delta + H^2 + c) . \]
Equality holds if and only if the mean curvature $H$ is constant (see [1], [5]). Moreover, the Killing spinor defines a map $f \mapsto \Phi (f)$ of the space $L^2 (M^2)$ of functions into the space $L^2 (M^2; S)$ of spinors such that
\[ ||D(\Phi (f)) ||^2_{L^2} = \langle \Delta f + H^2f+ c f,f \rangle_{L^2} . \]
In particular, the mentioned inequality holds for all eigenvalues, i.e.,
\[ \lambda_k^2 (D) \le \mu_k (\Delta + H + c) . \]
This inequality was used in order to estimate the first eigenvalue of the Dirac operator defined on special surfaces of Euclidean space (see [1]). On the other hand, in case we know $\lambda_1^2 (D)$, the inequality yields a lower bound for the spectrum of the Schr\"odinger operator $\Delta + H^2 + c$. For example, for any Riemannian metric $g$ on the 2-dimensional sphere $S^2$ the inequality
\[ \lambda_1^2 (D) \ge \frac{4 \pi}{\mbox{vol} \, (S^2,g)} \]
holds (see [2], [6]). Consequently, we obtain
\[ \frac{4 \pi}{\mbox{vol} \, (M^2,g)} \le \mu_1 (\Delta + H^2) \]
for any surface $M^2 \hookrightarrow {\Bbb R}^3$ of genus zero in Euclidean space ${\Bbb R}^3$. In this note we expose the described idea and, in particular, we estimate the spectrum of special periodic Schr\"odinger operators where the potential is given by the curvature $\kappa$ of a spherical curve. \\
\section{The one-dimensional case}
First of all, let us consider the 1-dimensional case, i.e., a curve $\gamma$ of length $L$ in a two-dimensional space form $X^2 (c)$. Let $\Phi$ be a Killing spinor of length one on $X^2 (c)$:
\[ \nabla_{T} {\Phi} = \frac{1}{2} c \cdot T \cdot \Phi . \]
The restriction $\varphi = \Phi_{|\gamma}$ defines a pair of spinors on $\gamma$ and the spinor field $\psi= f \cdot \varphi$ satisfies the equation:
\[ |D \psi |^2 = | \dot{f}|^2 +f^2 \left( \frac{c}{4} + \frac{1}{4} \kappa^2_g \right) , \]
where $\kappa_g$ is the curvature of the curve $\gamma$ in $X^2(c)$. Therefore, we obtain
\[ \lambda_k^2 (D) \le \mu_k \left( - \frac{d^2}{ds^2} + \frac{c}{4} + \frac{1}{4} \kappa^2_g \right) . \]
Suppose now that the spin structure on $\gamma$ induced by the spin structure of $X^2 (c)$ is non-trivial. Then we have $\displaystyle \lambda_{k+1}^2 (D)= \frac{4 \pi^2}{L^2} (k + 1/2)^2$ (see [4]) and, in particular, we obtain
\[ \frac{4 \pi^2}{L^2} \left( k + \frac{1}{2} \right)^2 \le \mu_{k+1} \left( - \frac{d^2}{ds^2} + \frac{c}{4} + \frac{1}{4} \kappa^2_g \right) . \]
{\bf Theorem 1:} {\it Let $\gamma \subset {\Bbb R}^3$ be a plane or spherical curve and denote by $\kappa^2 = c + \kappa_g^2$ the square of its curvature. Suppose that the induced spin structure on $\gamma$ is non-trivial, i.e., the tangent vector field has an odd rotation number. Then the inequality
\[ \frac{4 \pi^2}{L^2} \le \mu_1 \left( - 4 \frac{d^2}{ds^2} + \kappa^2 \right) \]
holds, where $\mu_1$ is the first eigenvalue of the periodic Sturm-Liouville operator on the interval $[0,L]$. Moreover, equality occurs if and only if the curvature is constant.}\\
{\bf Remark:} No geometric lower bound for the Sturm-Liouville operator $ - 4 \frac{d^2}{ds^2} + \kappa^2$ with potential defined by the square of the curvature $\kappa (s)$ of a closed curve $\gamma$ in Euclidean space seems to be known. We conjecture that the estimate given in \mbox{Theorem 1} holds for any closed curve in ${\Bbb R}^3$. Let us compare this inequality with the well-known Fenchel-Milnor inequality
\[ 2 \pi \le \oint\limits_{\gamma} \kappa . \]
Thus, by the Cauchy-Schwarz inequality we obtain
\[ \frac{4 \pi^2}{L^2} \le \frac{1}{L} \oint\limits_{\gamma} \kappa^2 . \]
Moreover, using the test function $f \equiv 1$, we have
\[ \mu_1 \left( - 4 \frac{d^2}{ds^2} + \kappa^2 \right) \le \frac{1}{L} \oint\limits_{\gamma} \kappa^2 . \]
Suppose that $\gamma$ is a simple curve in ${\Bbb R}^3$ and denote by $ \rho$ the minimal number of generators of the fundamental group $\pi_1 ({\Bbb R}^3 \backslash \gamma)$. Then we have
\[ 2 \pi \rho \le \oint\limits_{\gamma} \kappa . \]
In the spirit of this remark one should be able to prove the stronger inequality
\[ \frac{4 \pi^2}{L^2} \rho^2 \le \mu_1 \left( - 4 \frac{d^2}{ds^2} + \kappa^2 \right) \]
in case of a simple curve in ${\Bbb R}^3$. \\
{\bf Examples:} We calculated the eigenvalue $\mu_1$ for some classical curves in ${\Bbb R}^3$:
\begin{itemize}
\item[a.)] {\it The lemniscate:} $x= \sin (t) , \quad y = \cos (t) \sin (t).$\\
\mbox{} \hspace{2.7cm} $4 \pi^2 /L^2 = 1.06193, \quad \mu_1 = 3.7315, \quad \displaystyle \frac{1}{L} \oint\limits_{\gamma} \kappa^2 = 4.36004.$\\
\item[b.)] {\it The trefoil:} $x= \sin (3t) \cos (t) , \quad y = \sin (3t) \sin (t).$\\
\mbox{} \hspace{1.8cm} $4 \pi^2 /L^2 =0.221, \quad \mu_1 = 5.21, \quad \displaystyle \frac{1}{L} \oint\limits_{\gamma} \kappa^2 = 8.16.$\\
\item[c.)] {\it Viviani's curve:} $x= 1+ \cos (t) , \quad y = \sin (2t) , \quad z= 2 \sin (t) .$\\
\mbox{} \hspace{2.6cm} $4 \pi^2 /L^2 =0.169071, \quad \mu_1 = 0.5335, \quad \displaystyle \frac{1}{L} \oint\limits_{\gamma} \kappa^2 = 0.567803.$\\
\item[d.)] {\it Torus knot:} $x= (8+3 \cos (5t)) \cos (2t) , \quad y = (8+ 3 \cos (5t)) \sin (2t), \\
\mbox{} \hspace{2.3cm} z= 5 \sin (5t).$\\
\mbox{} \hspace{2.0cm} $4 \pi^2 /L^2 =0.00146034, \quad \mu_1 = 0.03232, \quad \displaystyle \frac{1}{L} \oint\limits_{\gamma} \kappa^2 = 0.0333803.$\\
\item[e.)] {\it The spherical spiral:} $x= \cos (t) \cos (4t) , \quad y = \cos (t) \sin(4t), \quad z= \sin(t).$\\
\mbox{} \hspace{3.4cm} $4 \pi^2 /L^2 =0.127036, \quad \mu_1 = 1.744, \quad \displaystyle \frac{1}{L} \oint\limits_{\gamma} \kappa^2 = 4.93147.$\\
\end{itemize}
\section{The two-dimensional Schr\"odinger operator}
We generalize this inequality to the case of the two-dimensional periodic Schr\"odinger operator
\[ P_{A,L} = - \left( 1+ \frac{A^2}{L^2} \right) \frac{\partial}{\partial t^2} - 4 \frac{\partial^2}{\partial s^2} - \frac{4 A}{L} \frac{\partial}{\partial t} \frac{\partial}{\partial s} + \kappa^2 (s) \]
defined on $[0,2\pi] \times [0,L]$:\\
{\bf Theorem 2:} {\it Let $\gamma \subset S^2 \subset {\Bbb R}^3$ be a closed, simple curve of length $L$ bounding a region of area $A$, and denote by $\kappa$ its curvature. Then the spectrum of the two-dimensional periodic Schr\"odinger operator $P_{A,L}$ is bounded by }
\[ \frac{4 \pi^2}{L^2} \le \mu_1 (P_{A,L}) . \]
{\it Equality holds if and only if the curvature of $\gamma$ is constant.}\\
In general, let us consider a Riemannian manifold $(Y^n,g)$ of dimension $n$ as well as an $S^1$-principal fibre bundle $\pi : P \to Y^n$ over $Y^n$. Denote by $\vec{V}$ the vertical vector field on $P$ induced by the action of the group $S^1$ on the total space $P$, i.e.,
\[ \vec{V} (p) = \frac{d}{dt} \left( p \cdot e^{it} \right)_{t=0} \quad , \quad p \in P . \]
A connection $Z$ in the bundle $P$ defines a decomposition of the tangent bundle $T(P)=T^v (P) \oplus T^h (P)$ into its vertical and horizontal subspace. We introduce a Riemannian metric $g^*$ on the total space $P$, requiring that
\begin{itemize}
\item[a)] $g^* (\vec{V} , \vec{V})=1$,\\
\item[b)] $ g^* (T^v, T^h)=0$,\\
\item[c)] the differential $d \pi$ maps $T^h (P)$ isometrically onto $T(Y^n)$.
\end{itemize}
\bigskip
A closed curve $\gamma : [0,L] \to Y^n$ of length $L$ defines a torus $H(\gamma) := \pi^{-1} (\gamma) \subset P$ and we want to study the isometry class of this flat torus in $P$. Let $\alpha = e^{i \Theta} \in S^1 $ be the holonomy of the connection $Z$ along the closed curve $\gamma$. Consider a horizontal lift $\hat{\gamma} :[0,L] \to P$ of the curve $\gamma$. Then
\[ \hat{\gamma} (L) = \hat{\gamma} (0) e^{i \Theta} \]
holds. Consequently, the formula
\[ \Phi (t,s) = \hat{\gamma} (s) e^{-i \Theta s / L } e^{it} \]
defines a parametrization $\Phi : [0,2 \pi] \times [0,L] \to H(\gamma)$ of the torus $H(\gamma)$. Since
\[ \frac{\partial \Phi}{\partial t} = \vec{V} \quad , \quad \frac{\partial \Phi}{\partial s} = dR_{e^{it} e^{-i \Theta s/ L}} (\dot{\hat{\gamma}} (s)) - \frac{\Theta}{L} \vec{V} , \]
we obtain
\[ g^* \left( \frac{\partial \phi}{\partial t} , \frac{\partial \phi}{\partial t} \right) = 1 , \quad g^* \left( \frac{\partial \phi}{\partial t} , \frac{\partial \phi}{\partial s} \right) = - \frac{\Theta}{L} , \quad g^* \left( \frac{\partial \phi}{\partial s} , \frac{\partial \phi}{\partial s} \right) = 1 + \frac{\Theta^2}{L^2} , \]
i.e., the torus $H(\gamma)$ is isometric to the flat torus $({\Bbb R}^2 / \Gamma_o , g^*)$, where $\Gamma_o$ is the orthogonal lattice $\Gamma_o = 2 \pi \cdot {\Bbb Z} \oplus L \cdot {\Bbb Z}$ and the metric $g^*$ has the non-diagonal form
\[ g^* = \left( \begin{array}{cc} 1 & - \frac{\Theta}{L} \\ \\- \frac{\Theta}{L} & 1+ \frac{\Theta^2}{L^2} \end{array} \right) . \]
Using the transformation
\[ x = - \frac{\Theta}{L} s + t \quad , \quad y=s , \]
we see that $H(\gamma)$ is isometric to the flat torus $({\Bbb R}^2 / \Gamma , \, \, \, dx^2 + dy^2 )$, where the lattice $\Gamma$ is generated by the two vectors
\[ v_1 = \left( \begin{array}{c} 2 \pi\\0 \end{array} \right) \quad , \quad v_2 = \left( \begin{array}{c} \Theta \\ L \end{array} \right) . \]
In case the closed curve $\gamma : [0,L] \to Y^n$ is the oriented boundary of an oriented compact surface $M^2 \subset Y^n$, we can calculate the holonomy $\alpha = e^{i \Theta}$ along the curve $\gamma$. Indeed, let $\Omega^Z$ be the curvature form of the connection $Z$. $\Omega^Z$ is a 2-form defined on the manifold $Y^n$ with values in the Lie algebra of the group $S^1$, i.e., with values in $i \cdot {\Bbb R}^1$. The parameter $\Theta$ is given by the integral
\[ \Theta = i \int\limits_{M^2} \Omega^Z . \]
Let us consider the Hopf fibration $\pi: S^3 \to S^2$, where
\[ S^3 = \{ (z_1, z_2) \in {\Bbb C}^2 : |z_1|^2 + |z_2|^2 =1 \} \]
is the 3-dimensional sphere of radius 1. The connection $Z$ is given by the formula
\[ Z= \frac{1}{2} \{ \bar{z}_1 dz_1 - z_1 d \bar{z}_1 + \bar{z}_2 dz_2 - z_2 d \bar{z}_2 \} \]
and its curvature form ($\omega = z_1 / z_2$)
\[ \Omega^Z = - \frac{d \omega \wedge d \bar{\omega}}{(1+|\omega|^2)^2} = - \frac{i}{2} dS^2 \]
essentially coincides with one half of the volume form of the unit sphere $S^2$ of radius 1. However, the differential $d \pi : T^h (S^3) \to T(S^2)$ multiplies the length of a vector by two, i.e., the Hopf fibration is a Riemannian submersion in the sense described before if we fix the metric of the sphere $S^2 (\frac{1}{2})= \{ x \in {\Bbb R}^3 : |x| = \frac{1}{2} \}$ on $S^2$. Consequently, in case of a closed simple curve $\gamma \subset S^2$ bounding a region of area $A$, the Hopf torus $H (\gamma) \subset S^3$ is isometric to the flat torus ${\Bbb R}^2 / \Gamma$ and the lattice $\Gamma$ is generated by the two vectors
\[ v_1 = \left( \begin{array}{c} 2 \pi\\0 \end{array} \right) \quad , \quad v_2 = \left( \begin{array}{c} {A}/{2} \\ {L}/{2} \end{array} \right) . \]
The mean curvature $H$ of the torus $H(\gamma) \subset S^3$ coincides with the geodesic curvature $\kappa_g$ of the curve $\gamma \subset S^2 \subset {\Bbb R}^3$ (see [7], [8]). We apply now the inequality
\[ \lambda_1^2 (D) \le \mu_1 ( \Delta + H^2 + 1) \]
to the Hopf torus $H (\gamma) \subset S^3$. Then we obtain the estimate
\[ \lambda_1^2 (D) \le \mu_1 \left( P_{A,L} \right) , \]
where $D$ is the Dirac operator on the flat torus ${\Bbb R}^2 / \Gamma$ with respect to the induced spin structure. All spin structures of a 2-dimensional torus are classified by pairs $(\varepsilon_1, \varepsilon_2)$ of numbers $\varepsilon_i =0,1$. If $\gamma$ is a simple curve in $S^2$, the induced spin structure on the Hopf torus $H(\gamma)$ is non-trivial and given by the pairs $(\varepsilon_1, \varepsilon_2) =(0,1)$. The spectrum of the Dirac operator for all flat tori is well-known (see [4]): The dual lattice $\Gamma^*$ is generated by
\[ v_1^* = \left( \begin{array}{c} \displaystyle \frac{1}{2 \pi} \\ \\ \displaystyle - \frac{A}{2 \pi L} \end{array} \right) \quad , \quad v_2^* = \left( \begin{array}{c} 0 \\ \\ \displaystyle \frac{2}{L} \end{array} \right) \]
and the eigenvalues of $D^2$ are given by
\begin{eqnarray*}
\lambda^2 (k,l) &=& 4 \pi^2 \Big| \Big| \, \, k v_1^* + \left(l + \frac{1}{2}\right) v_2^* \Big| \Big|^2 =\\
&=& k^2 + \frac{4 \pi^2}{L^2} \left((2l+1) - k \frac{A}{2 \pi} \right)^2 .
\end{eqnarray*}
We minimize $\lambda^2 (k,l)$ on the integral lattice ${\Bbb Z}^2$. The isoperimetric inequality $4 \pi A - A^2 \le L^2$ and $A \le \mbox{vol} \, (S^2) = 4 \pi$ yield the result that $\lambda^2 (k,l)$ attends its minimum at $(k,l) = (0,1)$, i.e.,
\[ \frac{4 \pi^2}{L^2} \le \lambda^2 (k,l) . \]
{\bf Remark 1:} We replace the Hopf fibration by the $S^1$-principal fibre bundle of Chern class $m \ge 0$. The corresponding total space is the Lens space $L(m,1)$ and we have the commutative diagram
\[
\begin{diagram}
\node{S^3} \arrow{se,b}{\pi} \arrow[2]{e} \node{} \node{L(m,1)} \arrow{sw,b}{\pi_m}\\
\node{} \node{S^2} \node{}
\end{diagram}
\]
Let $H_m (\gamma) \subset L(m,1)$ be the Hopf torus. $H_m (\gamma)$ is isometric to ${\Bbb R}^2 / \Gamma_m$, where the lattice $\Gamma_m$ is generated by the vectors
\[ v_1 = \left( \begin{array}{c} 2 \pi / m \\ 0 \end{array} \right) \quad , \quad v_2 = \left( \begin{array}{c} A/2 \\ L/2 \end{array} \right) . \]
Moreover, the Lens space $L(m,1)$ admits a unique spin structure with a Killing spinor (see [3]). Even in case of $m \not= 1$, the induced spin structure on $H_m (\gamma)$ is described by the parameters $(\varepsilon_1 , \varepsilon_2)=(0,1)$. Since the local geometry of $H_m (\gamma)$ in $L(m,1)$ essentially coincides with the geometry of $H(\gamma)$ in $S^3$, we obtain the inequality
\[ \frac{4 \pi^2}{L^2} = \min\limits_{(k,l)} \left\{ k^2m^2 + \frac{4 \pi^2}{L^2} \left( (2l+1) - k \frac{mA}{2 \pi} \right)^2 \right\} \le \mu_1 \left( - 4 \frac{d^2}{ds^2} + \kappa^2 \right) . \]
Consequently, the investigation of the two-dimensional Schr\"odinger operator in case of $m \not= 1$ yields the same result for the Sturm-Liouville operator as above. \\
{\bf Remark 2:} Suppose now that equality holds for some curve $\gamma \subset S^2$. We consider the corres\-ponding Hopf torus $H(\gamma) \subset S^3$ and then we obtain
\[ \lambda_1^2 (D) = \mu_1 ( \Delta + H^2 +1) . \]
Therefore, the mean curvature $H= \kappa$ is constant, i.e., $\gamma$ is a curve on $S^2$ of constant curvature $\kappa$. Consequently, $\gamma$ is a circle in a 2-dimensional plane. Denote by $r$ its radius. Then
\[ \kappa^2 = \frac{1}{r^2} \quad , \quad L= 2 \pi r \quad , \quad A=2 \pi (1 - \sqrt{1-r^2}) , \]
and the inequality
\[ \frac{4 \pi^2}{L^2} \le \kappa^2 \]
is an equality for all $r \not= 0$. \\
|
\section{Introduction}
If a body moves with a constant velocity, then, as is well known,
the body is relativistically contracted in the direction of motion,
whereas its length in the normal direction is unchanged. A naive
generalization to a rotating disc leads to the conclusion that the
circumference of the disc is contracted, whereas the radius of the
disc is unchanged. This paradox is known as the Ehrenfest paradox.
Obviously, the paradox is a consequence of the application of the
constant-velocity result to a system with a nonconstant velocity.
The standard resolution (see \cite{gron1}, \cite{gron2} and references
therein)
of the Ehrenfest paradox is as follows: One introduces the coordinates
of the rotating frame $S'$
\begin{equation}\label{eq1}
\varphi'=\varphi -\omega t \; , \;\;\;\; r'=r \; , \;\;\;\; z'=z
\; , \;\;\;\; t'=t \; ,
\end{equation}
where $\varphi$, $r$, $z$, $t$ are cylindrical coordinates of the
inertial frame $S$ and $\omega$ is the angular velocity. The metric in $S'$ is
given by
\begin{equation}\label{eq2}
ds^2=(c^2-\omega^2 r'^2)dt'^2 -2\omega r'^2 \, d\varphi' dt' -dr'^2
-r'^2 \, d\varphi'^2 -dz'^2 \; .
\end{equation}
It is generally accepted that the space line element should be
calculated by the formula \cite{land}
\begin{equation}\label{eq3}
dl'^2=\gamma'_{ij}dx'^i dx'^j \; , \;\;\; i,j=1,2,3 \; ,
\end{equation}
where
\begin{equation}\label{eq4}
\gamma'_{ij}=\frac{g'_{0i}g'_{0j}}{g'_{00}}-g'_{ij} \; .
\end{equation}
This leads to the circumference of the disc
\begin{equation}\label{eq5}
L'=\int_{0}^{2\pi} \frac{r' d\varphi'}{\sqrt{1-\omega^2 r'^2/c^2}}
=\frac{2\pi r'}{\sqrt{1-\omega^2 r'^2/c^2}}
\equiv\gamma(r')2\pi r' \; .
\end{equation}
The circumference of the same disc as seen from $S$ is $L=2\pi r=2\pi r'$.
If the disc is constrained to have the same radius $r$ as the same
disc when it does not rotate, then $L$ is not changed by the rotation, but
the proper circumference $L'$ is larger than the proper circumference of
the nonrotating disc. This implies that there are tensile stresses in the
rotating disc.
However, there is something wrong with this standard resolution
of the Ehrenfest paradox. Consider a slightly simpler situation;
a rotating ring in a rigid nonrotating
circular gutter with the radius $r=r'$.
The statement that (\ref{eq5}) represents the
proper circumference implies that the {\em proper} frame of the
rotating ring is given by (\ref{eq1}). This
means that an observer on the ring sees that the
circumference is $L'=\gamma L$. The circumference of the gutter
seen by him
cannot be different from the circumference of the ring
seen by him, so the
observer on the ring sees that the circumference of the
relatively moving gutter is {\em larger} than the proper
circumference of the gutter, whereas we expect that he should
see that it is smaller.
This leads to another paradox. It cannot be resolved by saying that
the observer on the ring accelerates, because one can consider a limit
$r\rightarrow\infty$, $\omega\rightarrow 0$, $r\omega\equiv u=$constant,
which implies that the acceleration $a=r\omega^2$ becomes zero,
whereas the paradox remains.
Before explaining how we resolve this paradox, we give some
general notes on the physical meaning of various coordinate frames in
the theory of relativity. In practice, one usually uses the
coordinates that simplify the technicalities of the
physical problem considered.
For example, when one describes physical effects in a
rigid body, it may be convenient to use a comoving coordinate frame,
i.e., a frame in which all particles of the rigid body have
constant spatial coordinates. The coordinates of $S'$ in
(\ref{eq1}) may be interpreted in this way.
However, {\em the choice of the
coordinate frame is more than a matter of convenience}.
The main lesson we have learned from Lorentz
coordinate frames is the fact that what an observer observes
(time intervals, space intervals, components of a tensor, ...)
depends on how the observer moves. The main purpose of
theoretical physics is to predict what will be {\em observed} under
given circumstances. Therefore, unless stated otherwise,
in this paper {\em by a
coordinate frame we understand a coordinate frame that
is inherent to an observer}, not to a set of physical particles.
Our criticism of some earlier treatments originates from
such an interpretation of coordinate frames. To avoid a possible
misunderstanding, we note that coordinate frames do not
necessarily need to be interpreted in this way, in which case
our criticism does not apply.
We resolve the paradox by recognizing that,
according to our interpretation, the frame defined by
(\ref{eq1}) is the proper frame {\em only} of the observer at
$r=r'=0$. This observer has no velocity relative to $S$,
so the corresponding coordinate transformation (\ref{eq1})
does not depend on any velocity. As will become clear from the discussion
of Section 2, the frame defined by (\ref{eq1}) is actually the Fermi
frame of an observer who rotates, but has no velocity with respect
to the frame $S$.
Observers at different positions
on the rotating disc have different velocities, so one has to use a
different coordinate transformation for each of them. In other
words, {\em although there is no relative motion among different points on a
rotating disc, each point belongs to a different noninertial
frame.}
This is not strange to those who
are familiar with the theory of Fermi coordinates \cite{mtw}, \cite{synge},
but it seems that many relativity-theorists are not.
Note also that since we do not interpret the coordinates of $S'$ in
(\ref{eq1}) as something inherent to the disc as a whole, $r'$ can be
arbitrarily large in (\ref{eq2}), although
there is a coordinate singularity at $r'=c/\omega$.
It resembles
the Schwarzschild singularity of a black hole, where the radial coordinate
is not restricted to be
larger than the Schwarzschild radius. However, to avoid a possible
misunderstanding, note that the coordinate singularity in (\ref{eq2}) does
not correspond to an event horizon, because a rotating observer at
$r'=0$ {\em can} receive information from $r'\geq c/\omega$.
There is also another paradox connected with the standard approach to
rotating frames. Let us consider how the nonrotating gutter looks like to
a rotating observer in the center. His proper frame
{\em is} given by (\ref{eq1}). If (\ref{eq3}) is the correct definition
of the space line element, then he should see that the circumference
of the gutter is larger
than the proper circumference of the gutter by a factor $\gamma(r')$.
However, $\omega r'/c$ can be
arbitrarily large, so $\gamma(r')$ can be not only
arbitrarily large, but also even imaginary. On the other hand, we know from
everyday experience that the apparent velocity $\omega r'$ of stars,
due to our rotation,
can exceed the velocity of light, but we see neither a contraction,
nor an elongation of the stars observed.
We resolve this paradox by examining the assumptions
under which formula (\ref{eq3}) is obtained. We find that this formula
should be used with great care and show that it is not applicable
in our case.
The correct definition of the space line element depends
on how it is measured, and we find that, in our case, $\gamma'_{ij}$ should be
replaced by $-g'_{ij}$ in (\ref{eq3}).
It is fair to note that there are also some other ``nonstandard"
approaches to the Ehrenfest paradox (see \cite{tart2}, \cite{kla},
and references
therein), but none of these approaches is similar to ours.
In particular, the crucial fact that
each point of the rotating ring belongs to a different frame has
not been taken into account in any of these approaches.
Formula (\ref{eq3}) has already been criticized \cite{kla}, but our criticism
of (\ref{eq3}) is quite different and more general.
The paper is organized as follows: In Section 2 we find the correct
coordinate transformation that leads to the
frame of an observer moving in flat spacetime.
In Section 3 we explain why (\ref{eq3})
is not always a correct definition of a space line element and
show that in a frame that corresponds to an observer in flat spacetime
it is more appropriate to calculate the space line element by
$-g'_{ij}$. We also make some general remarks on the physical meaning of
general coordinate transformations.
In Section 4 we study the relativistic contraction as seen by various
observers and resolve the Ehrenfest paradox. In Section 5 we study the rate
of clocks at various positions, as seen by various observers.
In Section 6 we discuss the velocity of light as seen by various observers.
In Section 7 we discuss our results, resolve some additional physical
problems, and give some generalizations. Section 8 is
devoted to concluding remarks, where the relevance of our results
to general relativity is emphasized.
\section{The frame of an observer moving in flat spacetime}
The generalized Lorentz transformations for a local Fermi
frame of an observer
that has arbitrary time-dependent velocity and angular velocity
in flat spacetime are found in
\cite{nels}. We present the final results, using
slightly different notation. Let $S$ be an inertial frame
and let $S'$ be the frame of the observer whose velocity and angular velocity
are $u^i(t')$ and $\omega^i(t')$, respectively, as seen by an observer in $S$.
The coordinate transformation between these two frames is given by
\begin{equation}\label{er1}
x^i =-A_{j}^{\; i}(t')x'^j +\int_{0}^{t'} \gamma(t')u^i(t')\, dt' +
\frac{1}{\mbox{\bf{u}}^2(t')} [\gamma(t')-1][u^k(t')A_{jk}(t')x'^j]u^i(t') \; ,
\end{equation}
\begin{equation}\label{er2}
t=\int_{0}^{t'} \gamma(t')\, dt' +
\frac{1}{c^2}\gamma(t')[u^k(t')A_{jk}(t')x'^j] \; ,
\end{equation}
where $\gamma(t')=1/\sqrt{1-\mbox{\bf{u}}^2(t')/c^2}$ and
$A_{ji}(t')=-A_{j}^{\; i}(t')$ is the
rotation matrix evaluated at $\mbox{\bf{x}}'=0$. The rotation matrix
satisfies the differential equation
\begin{equation}\label{er4}
\frac{d A_{ij}}{dt}=-A_{i}^{\; k}\omega_{kj} \; ,
\end{equation}
where $\omega_{ik}=\varepsilon_{ikl}\omega^{l}$, $\varepsilon_{123}=1$.
The metric tensor in $S'$ is
\begin{eqnarray}\label{metric}
& g'_{ij}=-\delta_{ij} \; , \;\;\;\;\;
g'_{0j}=-(\mbox{\boldmath $\omega$}'\times\mbox{\bf{x}}')_j \; , &
\nonumber \\
& g'_{00}=c^2 \left(
1+\displaystyle\frac{\mbox{\bf{a}}'\cdot\mbox{\bf{x}}'}{c^2}
\right)^2 -(\mbox{\boldmath $\omega$}'\times\mbox{\bf{x}}')^2 \; , &
\end{eqnarray}
where
\begin{equation}
\omega'^i =\gamma (\omega^i -\Omega^i) \; , \;\;\;\;\;
a'^i =\gamma^2 \left[a^i +\frac{1}{\mbox{\bf{u}}^2}(\gamma
-1)(\mbox{\bf{u}}\cdot\mbox{\bf{a}})u^i\right] \; ,
\end{equation}
$\Omega^{i}$ is the time-dependent Thomas precession frequency
\begin{equation}
\Omega_{i}=\frac{1}{2\mbox{\bf{u}}^2}(\gamma -1)\varepsilon_{ikj}
(u^k a^j -u^j a^k) \; ,
\end{equation}
and $a^i=du^i/dt$ is the time-dependent acceleration.
The transformations (\ref{er1})-(\ref{er2}) are chosen such that
the space origins of $S$ and $S'$ coincide for $t=t'=0$. If $\mbox{\bf{u}}$
is time independent and $\mbox{\boldmath $\omega$}=0$, then
(\ref{er1})-(\ref{er2}) reduce to the well-known ordinary Lorentz
boosts. If $\mbox{\bf{u}}=0$ and $\mbox{\boldmath $\omega$}$
is time independent, then (\ref{er1})-(\ref{er2}) reduce to (\ref{eq1}).
It is important to emphasize that $\mbox{\bf{u}}(t')$ is the velocity of the
{\em space origin} $\mbox{\bf{x}}'=0$ of $S'$. If $S'$ is a rotating frame, then
other space points of $S'$ have a different velocity. (Remind that rotation
is {\em not} a motion along a circle, but rather a change of orientation
of the axes with respect to an inertial frame.) Therefore,
in general, $S'$ is the proper frame {\em only}
of the observer at $\mbox{\bf{x}}'=0$.
Note also that $g'_{\mu\nu}=\eta_{\mu\nu}$ only at $\mbox{\bf{x}}'=0$,
which is another confirmation that $S'$ is the frame of the observer at
$\mbox{\bf{x}}'=0$ only. The metric (\ref{metric}) is also consistent
with a more general theory of Fermi coordinates \cite{mtw},
which are coordinates of an observer arbitrarily moving in curved spacetime,
and also have the property that $g_{\mu\nu}=\eta_{\mu\nu}$ at the
space origin, i.e., at the position of the observer. Note also that if
$\mbox{\bf{a}}'$ and $\mbox{\boldmath $\omega$}'$ vanish, then (\ref{metric})
is a metric of an inertial frame and is equal to $\eta_{\mu\nu}$
everywhere, so, in this case, $S'$ can be considered as a frame of an observer at
{\em arbitrary} constant $\mbox{\bf{x}}'$.
It is interesting to note that the geometrical
construction of Fermi coordinates is well established \cite{mtw},
\cite{synge}, but no analog of (\ref{er1})-(\ref{er2}) is
known for curved spacetime. The transformations (\ref{er1})-(\ref{er2})
are obtained by summation of infinitesimal Lorentz transformations (and
rotations). It is not so easy to find an analog of Lorentz transformations
in curved spacetime, because they correspond to the coordinate
transformation between Fermi frames of two different free-falling observers.
We can, however,
write the transformations (\ref{er1})-(\ref{er2})
in a more elegant form, which could be illuminating
for a generalization to curved spacetime. Let
\begin{equation}\label{el1}
x^{\mu}=f^{\mu}(t',\mbox{\bf{x}}';\mbox{\bf{u}})
\end{equation}
denote the ordinary Lorentz transformations, i.e., the transformations
between two inertial frames specified by the relative velocity
$\mbox{\bf{u}}$, which can be considered as the relative velocity
between two inertial (free-falling)
observers at the instant when they have the same position. The differential
of (\ref{el1}) is
\begin{equation}\label{el2}
dx^{\mu}=f^{\mu}_{\; ,\nu} (t',\mbox{\bf{x}}';\mbox{\bf{u}}) dx^{\nu} \; .
\end{equation}
The transition to a noninertial frame introduces a time-dependent
velocity: $\mbox{\bf{u}} \rightarrow \mbox{\bf{u}}(t')$.
The transformations (\ref{er1})-(\ref{er2}) may be obtained by integrating
(\ref{el2}) in the following way:
\begin{equation}\label{el3}
x^{\mu}=\int_{0}^{t'}f^{\mu}_{\; ,0} (t',0;\mbox{\bf{u}}(t')) dt' +
\int_{C}
f^{\mu}_{\; ,i} (t',\mbox{\bf{x}}';\mbox{\bf{u}}(t')) dx'^{i} \; ,
\end{equation}
where $C$ is an arbitrary curve with constant $t'$, starting from $0$
and ending at $-A_{j}^{\; i}(t')x'^{j}$.
The subintegral function in the second term
of (\ref{el3}) is a total derivative, so this term does not depend on the
curve $C$ and can be easily integrated. The time derivative in the
first term is taken with $\mbox{\bf{u}}(t')$ kept fixed, so
$f^{\mu}_{\; ,0}$ in this term is not a total derivative.
Let us now apply the general formalism described in this section to a
uniformly rotating ring. We assume that the ring is put in a rigid
nonrotating circular gutter with the radius $R$,
which provides that the radius of the rotating ring
is the same as the radius of the same ring when it does not rotate,
and is equal to $R$,
as seen by an observer in $S$. This allows us not to worry about
the complicated dynamical forces that tend to change the radius of the
ring as seen by the observer in $S$, and pay all our attention to the
kinematic effects resulting from the transformations
(\ref{er1})-(\ref{er2}).
The ring can be considered as a series of independent short rods,
uniformly distributed along the gutter. (By a short rod we
understand a rod with a length much shorter than $R$.) We assume
that the gutter is placed at the $z=0$ plane. We put the
space origin of $S$ at a fixed point on the gutter, such that
the $y$-axis is tangential to the gutter and the $x$-axis is perpendicular
to the gutter at $\mbox{\bf{x}}=0$. (In the rest of this section, as well as
in Sections 4 and 5,
$\mbox{\bf{x}}\equiv
( x,y)$ and the $z$-coordinate is suppressed.)
We study a single short rod initially placed at $\mbox{\bf{x}}=0$ and
uniformly moving along the gutter in the counterclockwise direction.
(This mimics a uniform motion of an electron in a
synchrotron). The gutter causes a torque that provides that the rod is
always directed tangentially to the gutter. Therefore,
$\omega=u/R$, where $u=\sqrt{\mbox{\bf{u}}^2}$ is time independent.
Now, $\gamma=1/\sqrt{1-\omega^2 R^2/c^2}$ is also time independent.
Since a clock in $S'$ is at $\mbox{\bf{x}}'=0$, the clock rate between
a clock in $S$ and a clock in $S'$ is given by $t=\gamma t'$, as seen
by an observer in $S$. We assume that, initially, the axes $x'$, $y'$ are
parallel to the axes $x$, $y$, respectively. Therefore the velocity
\begin{equation}
\mbox{\bf{u}}(t')=\omega R (-\sin \gamma\omega t', \cos \gamma\omega t')
\end{equation}
is always in the $y'$-direction and the solution of (\ref{er4}) is
\begin{equation}
A_{ij}(t')=\left(
\begin{array}{cc}
\cos \gamma\omega t' & \sin \gamma\omega t' \\
-\sin \gamma\omega t' & \cos \gamma\omega t'
\end{array} \right) \; .
\end{equation}
The transformations (\ref{er1})-(\ref{er2}) become
\begin{equation}\label{er1'}
\left( \begin{array}{c} x \\
y
\end{array} \right)=
\left( \begin{array}{cc}
\cos \gamma\omega t' & -\gamma\sin \gamma\omega t' \\
\sin \gamma\omega t' & \gamma\cos \gamma\omega t'
\end{array} \right)
\left( \begin{array}{c} x' \\
y'
\end{array} \right)
+R \left( \begin{array}{c} \cos \gamma\omega t' -1 \\
\sin \gamma\omega t'
\end{array} \right) \; ,
\end{equation}
\begin{equation}\label{er2'}
t=\gamma t' +\frac{\gamma}{c^2}\omega R y' \; .
\end{equation}
In particular, at $t'=0$ these transformations become
\begin{equation}\label{t=0}
x=x' \; , \;\;\;\; y=\gamma y' \; , \;\;\;\; t=\frac{\gamma u}{c^2}y' \; ,
\end{equation}
which coincide with the ordinary Lorentz boost at $t'=0$ for the velocity in
the $y$-direction.
\section{General coordinate transformations and the space
line element in a non-time-orthogonal frame}
A non-time-orthogonal frame is a frame in which $g'_{0i}$ is different
from zero. It is generally accepted that the space line element in such a frame
is given by (\ref{eq3}). However, if we assume that this formula can be
applied to calculate the space distance as seen by a local observer,
then, as we have found in Section 1, Eq. (\ref{eq3})
leads to an imaginary length of a distant unaccelerated object as seen
by a rotating observer. In order to resolve this puzzle, we examine the
assumptions under which formula (\ref{eq3}) is derived.
In \cite{land}, formula (\ref{eq3}) is derived by assuming that the space
distance between two points is measured by measuring the time $\Delta t'$ that
light needs to travel from point $A$ to point $B$ and then back to
point $A$. It is also assumed that the time is measured by a clock
that does not change its position $x'^i$.
The definition of the space distance $l'=c\,\Delta t'/2$ leads to
(\ref{eq3}).
In order to perform the described measurement in a
rotating frame, the clock must be
positioned at point $A$. However, according to our interpretation of
(\ref{eq1}), this point can be faraway
from the center of the rotation, so the required velocity
of point $A$ can exceed $c$, as seen in $S$. Therefore,
in general, such a measurement cannot be performed.
In practice, we measure space distances between distant objects in a
completely different way, namely, by measuring the angles under which we see the
objects. (We assume that we know the radial distance of these objects from
us. The radial distance is not problematic in the theoretical sense, because
$g'_{0r}=0$ in (\ref{eq2})).
Our rotation does not influence this angle. Therefore,
the apparent velocity of distant objects
can exceed the velocity of light owing to our rotation,
but a pure rotation (without velocity)
will not lead to relativistic contraction, nor to elongation. The effect is
that, in a rotating frame, it is more appropriate to calculate the space
line element as
\begin{equation}\label{dl}
dl'^2=-g'_{ij}dx'^{i}dx'^{j} \; ,
\end{equation}
despite the fact
that $g'_{0i}$ is different from zero. This formula should be used
to calculate the space distance between two arbitrary points
which have the same $t'$ coordinate, no matter
how far these points are from the observer at $x'^i =0$.
Of course, if these points are end points of a body,
then, in general, the distance calculated in this way
will not be equal to the proper length of the body, but merely to
the length seen by the observer.
Formula (\ref{dl})
is also correct for
frames that are both accelerated and rotating, defined by
(\ref{er1})-(\ref{er2}).
To clarify the meaning of formula (\ref{eq3}) completely, note that in
\cite{mol} this formula is derived in a completely different way,
without referring to any particular method of measurement. However,
what is actually derived in \cite{mol} is the fact that the
quantity (\ref{eq3}) does not change under coordinate transformations
of the form
\begin{equation}\label{sameframe}
t''=f^0 (t',x'^1,x'^2,x'^3) \; , \;\;\;\; x''^i=f^i (x'^1,x'^2,x'^3) \; .
\end{equation}
We refer to such transformations as {\em internal transformations}.
Obviously, (\ref{eq1}) is not an internal transformation.
Regular internal transformations form a subgroup of the group of all
regular coordinate transformations.
Note that the invariant quantity
$ds^2=g_{\mu\nu}dx^{\mu}dx^{\nu}$ can always be written as
\begin{equation}\label{logun}
ds^2=d\eta^2 -\gamma_{ij}dx^i dx^j \; ,
\end{equation}
where
\begin{equation}\label{logun2}
d\eta^2 =\left[ \frac{g_{0\mu}dx^{\mu}}{\sqrt{g_{00}}} \right]^2 \; ,
\end{equation}
so $d\eta^2$ also does not change under internal
transformations. The quantity $d\eta^2$ is nothing else but a time line element
\cite{land}, defined by a measuring procedure similar to
the measuring procedure used to define
the space line element (\ref{eq3}).
Let us illustrate the power of (\ref{eq3}), (\ref{sameframe}), and
(\ref{logun2}) on the example that has already been discussed at some
length in \cite{mol}. The Galilei transformation
$t''=t$, $x''=x-ut$ can also serve as a correct coordinate transformation
needed
to describe the relativistic effects related to a frame moving with a constant
velocity $u$. The metric in these coordinates is given by
\begin{equation}\label{mol1}
ds^2=c^2(1-u^2/c^2)dt''^2 -2udx''dt''-dx''^2 \; ,
\end{equation}
where it has been assumed that the metric of $S$ is given by
$ds^2=c^2\, dt^2 -dx^2$. From (\ref{eq3}) and $dt=0$ one can
obtain the relativistic contraction $dl=dx=dl''/\gamma$,
where $\gamma =1/\sqrt{1-u^2/c^2}$. Similarly,
from (\ref{logun2}) and $dx''=0$ one can obtain $dt=\gamma d\eta''$.
The frame $S''$ is physically equivalent to the frame
$S'$ which would be obtained from $S$ by the ordinary Lorentz
transformations, in the sense that $S''$ and $S'$ are connected by an
internal coordinate transformation
\begin{equation}\label{mol2}
x'=\gamma x'' \; , \;\;\;\;\; t'=t''/\gamma -\gamma u x''/c^2 \; .
\end{equation}
Note that the non-time-orthogonal metric (\ref{mol1}), unlike (\ref{eq2}) and
(\ref{metric}), can be transformed to a time-orthogonal metric by an
{\em internal} transformation. Note also that the metric (\ref{mol1}), unlike
(\ref{eq2}) and (\ref{metric}), is not a metric of a Fermi frame.
In \cite{mol}, internal transformations are interpreted as transformations
that correspond to a redefinition of the coordinates of the same {\em
physical} observer. However,
there is something unphysical about internal transformations; if $t'$ is a measure
of the physical time for the observer in $S'$, then $t''$ is not, because
it corresponds to a ``time" of the same observer which depends on the
space point
$x'^i$. Therefore, we introduce a more restrictive class of coordinate
transformations, which could be better suited to interpret them
as transformations that correspond to a redefinition of the coordinates of
the same physical observer:
\begin{equation}\label{veryweak}
t''=f^0 (t') \; , \;\;\;\; x''^i=f^i (x'^1,x'^2,x'^3) \; .
\end{equation}
We refer to such transformations as {\em restricted internal transformations}.
Regular restricted internal transformations form a subgroup of the group of all
regular internal transformations.
The quantities $g'_{00}dt'^2$ and (\ref{dl}) do not change under
restricted internal transformations.
Now we have two definitions of the space line element,
(\ref{eq3}) and (\ref{dl}), and related to this, two types
of restricted coordinate transformations, internal and restricted internal.
The space line element (\ref{eq3}) reduces to (\ref{dl}) if
$g_{0i}=0$. However, as we have shown in this section,
(\ref{dl}) is more appropriate in some cases,
even if $g_{0i}\neq 0$. How to
know in general what is the suitable definition of the
space line element?
We can immediately formulate one rule which is certainly
suitable: {\em If the metric of a frame can be transformed to a
time-orthogonal frame by an internal transformation, then the
space line element should be calculated by (\ref{eq3})}.
According to the results of this section, we can also formulate
another rule: {\em If the metric of a frame in flat spacetime can be
obtained from $g_{\mu\nu}=\eta_{\mu\nu}$ by a transformation of the form of
(\ref{er1})-(\ref{er2}) followed by an arbitrary restricted internal transformation, then
the space line element should be calculated by (\ref{dl}).}
Such coordinate transformations can be interpreted as the most general coordinate
transformations in flat spacetime that correspond to a physical observer
who has a positive mass.
We still do not know a general rule. However, one can be satisfied
to have a rule for Fermi frames only, or for Fermi frames modified
by an arbitrary restricted internal transformation, because only such frames have a direct
physical interpretation. One can be tempted to guess that
for all such frames the space line element should be calculated by
(\ref{dl}), but such a conjecture requires further investigation.
For the sake of completeness, let us make a few remarks on general
coordinate transformations in curved spacetime.
The most general coordinate transformation
that corresponds to a physical observer who has a positive mass is a
transformation that leads
to Fermi coordinates, followed by an arbitrary restricted internal transformation.
Other coordinate frames may be useful for some physical calculations,
for example, because it is easier to solve some covariant equations of motion
in these coordinates. However, if one is interested in how
the physical system
looks like to a physical observer, one must transform the results to the
coordinates specific for this observer.
To summarize this section, we conclude that the correct definition
of the space line element depends on how it is measured. Formula
(\ref{eq3}) is not incorrect, but its applicability is limited and
it should be used with great care. In our case of
accelerated, rotating frames,
it is more appropriate to calculate the space line element with
$-g'_{ij}$ instead of with $\gamma'_{ij}$.
\section{Relativistic contraction}
In Section 2 we have found the coordinate transformation that describes
the frame of a short rod uniformly moving along the circular gutter.
Let as assume for a while that the length of
the rod is infinitesimally small and
that the rod is rigid (i.e., its proper length $dL'$ is equal to the
proper length of the same rod when it does not accelerate).
Let us determine the relativistic contraction of the rod, as seen by
an observer in $S$. The observer in $S$ sees both ends of the rod
at the same instant, so $dt=0$.
From symmetry it is obvious that the relativistic contraction cannot depend on
$t$, so, in order to simplify the calculations, we evaluate this at $t=0$.
Since the rod is at $x'=y'=0$, (\ref{er2'}) implies that $t'=0$. Taking the
differential of (\ref{er1'}) and (\ref{er2'}) with respect to space and
time coordinates, and then putting $x'=y'=t'=dt=0$, we find that the
observer in $S$ sees the length
\begin{equation}\label{inf}
dL=dy=\frac{dy'}{\gamma}=\frac{dL'}{\gamma} \; ,
\end{equation}
which is the expected relativistic contraction.
Let us now turn our attention to the concept of the proper length of a
body. Traditionally, it is defined as a length of the body as seen
from the proper frame of the body. However, as we have seen,
in general, there is no
such thing as a proper frame of the body as a whole. Such a thing
exists only for a nonrotating, inertially moving body in flat
spacetime. The concept of a proper length
of a large body does not have any fundamental meaning,
simply because a ``large body" is not actually one object, but a
set of many interacting particles. However,
the proper length of an infinitesimally small part of a body is well
defined. Therefore, we can define the proper length of a whole
body as the sum of the proper lengths of its infinitesimal parts.
Applying this to (\ref{inf}), we see
that the relativistic contraction of a short (but not infinitesimal)
rigid rod uniformly moving along the circular gutter
is given by $L=L_0/\gamma$, as seen by the
observer in $S$. Here
$L_0$ is the proper length defined as above.
Now, as in Section 2, assume that the rotating ring is a series of
independent short rods, uniformly distributed along the gutter.
Each rod is relativistically contracted, but the ring is not.
This means that the
distances between the neighboring ends of the neighboring rods
are larger than those for a nonrotating ring, so
the proper length of the ring is also larger than that of a nonrotating
ring. This is concluded
also in \cite{gron2}. This situation mimics a more realistic
ring made of elastic material, where atoms play the role of
short rigid rods. Owing to the rotation the distances between
neighboring atoms
increase, so there are tensile stresses in the material.
However, it is important to emphasize that the rotation is not
essential for understanding of the origin of these tensile forces,
because a similar effect also occurs in a linear relativistic
motion \cite{dew}.
The same relativistic contraction of
short rods will be seen by a
rotating observer in the center, because his frame is given by the
Galilei transformation (\ref{eq1}) and the lengths are calculated by
$g_{ij}$, as explained in Section 3.
Let us now study how the nonrotating gutter looks like
from the point of view of an observer on the rotating ring.
Without
losing on generality, we evaluate this at $t'=0$. We calculate the
length of an infinitesimal part of the gutter lying near the
observer, so $x=y=0$. Both ends are seen at the same instant, so
$dt'=0$. Taking the
differential of (\ref{er1'}) with respect to space coordinates,
and then putting $t'=0$, we find that the
observer in $S'$ sees the length
\begin{equation}\label{inf2}
dL'=dy'=\frac{dy}{\gamma}=\frac{dL}{\gamma} \; ,
\end{equation}
which is the expected relativistic contraction.
It is important to emphasize that (\ref{inf2}) is correct only in the
infinitesimal form. The observer on the ring will not see other distant
parts of the gutter contracted in the same way; for him, the gutter and
the ring do not look azimuthally symmetric.
In the following
we study how other parts of the ring look like from the point of view
of the observer on the ring. We introduce polar
coordinates $r$, $\varphi$, defined by
\begin{equation}
y=r \sin \varphi \; , \;\;\;\; R+x=r \cos \varphi \; ,
\end{equation}
which are new space coordinates for $S$, with the origin in the center
of the circular gutter. The angle $\varphi$ is a good label of the position
of any part of the ring even in $S'$. (To visualize this, one can draw
angular marks on the gutter. The number of marks separating two points
on the gutter or on the ring is a measure of the ``angular distance" in
any frame.)
Let $S''$ be the frame of another
part of the ring. The position of that part of the ring is
$x''=y''=0$. The relative position of the space origin of $S''$ with respect
to that of $S'$ is given by the constant relative angle $\Delta\varphi_0$,
as seen by an observer in $S$. In analogy
with (\ref{er1'})-(\ref{er2'}), we find that $S''$ is determined by
\begin{equation}\label{er1''}
\left( \begin{array}{c} x \\
y
\end{array} \right)=
\left( \begin{array}{cc}
\cos (\gamma\omega t''+\Delta\varphi_0) & -\gamma\sin (\gamma\omega
t''+\Delta\varphi_0) \\
\sin (\gamma\omega t''+\Delta\varphi_0) & \gamma\cos (\gamma\omega
t''+\Delta\varphi_0)
\end{array} \right)
\left( \begin{array}{c} x'' \\
y''
\end{array} \right)
+R \left( \begin{array}{c} \cos (\gamma\omega t''+\Delta\varphi_0) -1 \\
\sin (\gamma\omega t''+\Delta\varphi_0)
\end{array} \right) \; ,
\end{equation}
\begin{equation}\label{er2''}
t=\gamma t'' +\frac{\gamma}{c^2}\omega R y'' \; .
\end{equation}
The observer in $S'$ will see the other part of the ring at the
relative ``angular distance" $\Delta\varphi$, which, owing to the relativistic
effects, differs from $\Delta\varphi_0$. Let the labels $A$, $B$ denote the
coordinates of the part of the ring that lie at $S'$ and $S''$,
respectively. Since the rotation is uniform,
the relative ``angular distance"
\begin{equation}\label{Dphi}
\Delta\varphi=\varphi_{B}(t''_{B})-\varphi_{A}(t''_{A})
= \Delta\varphi_0 +\gamma\omega t''_{B}-\gamma\omega t'_{A} \; ,
\end{equation}
cannot depend on $t'$, so without losing on generality, we evaluate this at
$t'=0$. Since the observer sees both parts of the ring at the same
instant, we have $t'_{A}=t'_{B}=0$. Since $x''_{B}=y''_{B}=0$, from
(\ref{er1''}) we find
\begin{equation}\label{y2}
y_{B}=R\sin (\gamma\omega t''_{B}+\Delta\varphi_0) \; ,
\end{equation}
and from (\ref{er2''})
\begin{equation}\label{t2}
t_{B}=\gamma t''_{B} \; .
\end{equation}
From $t'_{B}=0$ and (\ref{t=0})
it follows $t_{B}=\omega R y_{B}/c^2$, which, because
of (\ref{t2}), can be written as $\gamma t''_{B}=\omega R y_{B}/c^2$.
This, together with (\ref{y2}), leads to the equation that determines
$t''_{B}$:
\begin{equation}\label{eqgron1}
\gamma\omega t''_{B}=\beta^2\sin (\gamma\omega
t''_{B}+\Delta\varphi_0) \; ,
\end{equation}
where $\beta^2 \equiv \omega^2 R^2 /c^2$.
From $t'_{A}=0$ and (\ref{Dphi}) we see that $\Delta\varphi=\gamma\omega
t''_{B}+\Delta\varphi_0$, so (\ref{eqgron1}) can be written as
\begin{equation}\label{eqgron1'}
\Delta\varphi -\Delta\varphi_0 =\beta^2\sin \Delta\varphi \; .
\end{equation}
Equation (\ref{eqgron1'}) determines the relative ``angular
distance" $\Delta\varphi$
between two points on the ring as seen by the observer at one of the
points, if the relative angle between these two points, as seen by the
observer in $S$, is $\Delta\varphi_0$. In other words, (\ref{eqgron1'}) determines
how the ring looks like to the observer on the ring. For an inertial
observer whose instantaneous
position and velocity are equal to that of the observer on the ring,
the same equation (\ref{eqgron1'}) is found in \cite{gron1}, where
the solution is graphically depicted. This means, contrary to the
conclusion of \cite{gron1}, that the inertial and the noninertial
observers see the ring in the same way.
If the two points on the ring are very close to each other, then
$\Delta\varphi_0$ and $\Delta\varphi$ are very small. By expanding
equation (\ref{eqgron1'}) for small angles we find the approximative
solution $\Delta\varphi=\gamma^2 \Delta\varphi_0$. The factor $\gamma^2$ is
easy to understand; one factor of $\gamma$ appears because the
part of the gutter close to the observer on the ring looks shorter
for that observer than it really is, and the other factor of $\gamma$
appears because the part of the ring close to the observer on the ring
is longer than that of the same ring when it does not rotate.
\section{The rate of clocks}
Assume that there are two clocks at different positions on the
ring. Assume also that they show the same time, as seen by an
observer in $S$. Then, as shown in Section 2, both clocks
show the time $t'=t/\gamma$, as seen from $S$.
These two clocks do not show the same time as seen by an observer
on the ring. If the position of the observer coincides with the
position of one of the clocks, then the time-shift of the other
clock is given by (\ref{eqgron1}).
Let us calculate the time-shift of the clock at the fixed position
$(x,y)$, as seen by the observer in $S'$. From (\ref{er1'})
we express $y'$ as a function of $x$, $y$, and $t'$, and put this
in (\ref{er2'}). The result is
\begin{equation}\label{txy}
t=\gamma t' +
\frac{\omega R}{c^2}[y \cos \gamma \omega t'
-(x+R) \sin \gamma \omega t'] \; .
\end{equation}
For comparison, if (\ref{er1'}) and (\ref{er2'}) are replaced by
the ordinary Lorentz boosts for a constant velocity in the $y$-direction,
then (\ref{txy}) should be replaced by
\begin{equation}\label{txyo}
t=\frac{t'}{\gamma}+\frac{u}{c^2}y \; .
\end{equation}
To understand the physical meaning of (\ref{txy}),
we explore some special cases.
If $\gamma \omega t'=2k\pi$, then $t=\gamma t'+\omega R y/c^2$. In this
case, the rate of clocks $\Delta t/\Delta t'=\gamma$ is the same as
that for the observer
in $S$. This can also be understood as a time-averaged rate,
because the oscillatory functions in (\ref{txy}) vanish when
they are averaged over time. Therefore, the observer in $S'$ agrees
with the observer in $S$ that the clock in $S'$ is slower, but only in
a time-averaged sense.
At some instants the observer in $S'$ sees that
the clock in $S$ is slower than his clock.
For example, by putting $x=0$ and
expanding (\ref{txy}) for small $t'$, we recover formula (\ref{txyo}),
with $u=\omega R$. If the clock in $S$ is in the center, which corresponds
to $x=-R$, $y=0$, then (\ref{txy}) gives $t=\gamma t'$, so in this case
there is no oscillatory behavior.
\section{Velocity of light}
Let us also make some comments on the velocity of light. The Sagnac effect
is usually interpreted as a dependence of the velocity of light
on the direction of light propagation in a rotating frame
(see, for example, \cite{post}, \cite{tart} and references therein).
However, such an interpretation is based on the interpretation
of the frame $S'$ defined by (\ref{eq1}) as a proper frame of all
observers on a rotating platform. Now we know that each observer
belongs to a different local Fermi frame, and from (\ref{metric}) we see that
in the {\em vicinity} of any observer the metric is equal to the
Minkowski metric $\eta_{\mu\nu}$. This implies that {\em
for any local observer the velocity of light is isotropic and is
equal to $c$, providing that it is measured by propagating a light
beam in a {\bf small} neighborhood of the observer,
using Einstein synchronized clocks}. This is
also true for an observer in curved spacetime, because his
proper frame is given by the appropriate Fermi coordinates, which
also have a property that $g_{\mu\nu}=\eta_{\mu\nu}$ at the position
of the observer. The phrases
``local" and ``small" denote spatial dimensions inside which
the metric tensor does not change significantly.
Of course, the velocity
of light does not have to be equal to $c$ for an observer which is not
at the same position as the light. However, this is not only a property
of non-time-orthogonal frames. For example, if the acceleration
of an uniformly accelerated observer and the propagation of light
are both in the $x'$-direction, then from (\ref{metric}) one can
find that the accelerated observer sees the velocity of light as
$|dx'/dt'|=c\sqrt{1+a'x'/c^2}$, being equal to $c$ only at
$x'=0$. A similar effect occurs for a radial
motion of light in the vicinity of the Schwarzschild radius of a black hole,
as seen by a static observer faraway from the Schwarzschild radius.
Concerning the Sagnac effect, we do not claim that the standard prediction
for the phase shift is incorrect. It can also be derived by performing
calculations in the nonrotating frame $S$ \cite{post}, and such a
derivation, based on the well-understood Minkowski
spacetime, is perfectly correct. We have nothing new to say about
the phase shift, which appears when clockwise and counterclockwise
propagated
light beams finally meet. However, as seen by an observer on the rim of a
rotating disc, the velocity of the light beam will be a complicated function
of time $t'$, or equivalently, of the position $(x',y')$ of the beam.
The trajectory of
the light beam expressed in $S$-coordinates takes a simple form
\begin{equation}\label{sag}
y=R \sin \omega_L t \; , \;\;\;\;\;
x=R(-1+\cos \omega_L t) \; ,
\end{equation}
where $\omega_L =\pm c/R$. The plus and minus signs refer to the
counterclockwise and clockwise propagated beams, respectively.
Using (\ref{er1'}), (\ref{er2'}), and (\ref{sag}), one can eliminate
$x,y,t$ and express $x',y'$ as functions of $t'$. The speed of
light as seen by the observer in $S'$ is
\begin{equation}\label{sag2}
v'_L =\sqrt{\left( \frac{dx'}{dt'}\right)^2 +
\left( \frac{dy'}{dt'}\right)^2 } \; .
\end{equation}
Expanding (\ref{er1'}) and (\ref{sag}) for small $t'$ and $t$, respectively,
one can easily find $y'=\pm ct'+{\cal O}(t'^2)$, $x'={\cal O}(t'^2)$,
which means that the observer sees the velocity
of light equal to $c$ when the light is at the same position
as the observer, just as expected.
\section{Discussion}
From the experience acquired by careful calculations in the preceding
sections, we can generalize some of the results without much effort,
using qualitative and intuitive arguments.
If an observation in $S$ is performed at the instant $t$, then the
solution of (\ref{er4}) can always be chosen such that at $t$
the axes $x'^i$ are parallel to the corresponding axes $x^i$. Therefore,
for a small range of values of $t'$, the transformations
(\ref{er1})-(\ref{er2}) can be approximated by the ordinary Lorentz boosts (see
(\ref{t=0})).
From this fact we conclude that
if a moving rigid body is short enough, then its relativistic contraction
in the direction of the instantaneous velocity, as seen from $S$,
is simply given by $L(t)=L'/\gamma(t)$, i.e., it depends
only on the instantaneous velocity, not on its acceleration and rotation.
(``Short enough" means that $L'\ll c^2/a'_{\|}$, where $a'_{\|}$ is the
component of the proper acceleration parallel to the direction
of the velocity \cite{nikol}).
By a similar argument we may conclude that
an arbitrarily accelerated and rotating
observer sees equal lengths of other differently
moving objects as an inertial observer whose
instantaneous position and velocity are equal to that of the
arbitrarily accelerated and rotating observer.
So far we have studied a rotating ring. A rotating disc is a more
complicated object, with some additional dynamical effects
related to elastic and inertial forces. However, a disc can be modeled
as a series of concentric rings, each of them being constrained to have a
fixed radius. In this case, the analysis of a rotating disc becomes
essentially the same as that of a rotating ring.
Let us also give some additional arguments why our resolution of the
Ehrenfest paradox is correct. Our method, based
on coordinate transformations (\ref{er1})-(\ref{er2}), is really
a generalization of the well-known derivation of the Lorentz contraction
for constant velocities. In our
approach the origin of the relativistic contraction lies in
the non-Galilean transformation, not in the nontrivial metric, whereas
in the standard approach the transformation is Galilean and
the contraction is due to the nontrivial metric (\ref{eq2}).
Note finally that our approach allows a generalization to a more
complicated motion, whereas the standard approach does not.
Finally, let us make some comments on the observability of the relativistic
contraction. In principle, it could be observed by photographing a rod
with a very short exposition, such that both ends are observed
at the same instant. Since the velocity of the incoming information
(velocity of light) is finite, both ends of the rod should be
positioned at the same distance from the observer. Therefore, the ideal
setup for such a measurement is a rod in a uniform circular
motion and a camera in the center, providing that we can achieve a
short enough exposition. It is assumed that in this experiment the
only object that moves circularly is a rod (with two ends); there is
neither a rotating disc, nor a rotating ring.
An indirect, but easier-to-perform experimental
verification of the relativistic contraction could perhaps be obtained
by measuring the velocity of a rotating ring in a rigid circular
gutter, needed to achieve the break of the ring, and comparing it with
the elongation needed to achieve the break of the ring
caused by ordinary stretching.
Of course, in both types of experiments
the problem is to achieve a relativistic velocity
of macroscopic objects, so these can be
considered merely as {\it gedanken} experiments.
\section{Conclusion}
In this paper a new resolution of the Ehrenfest paradox has been
provided by taking into consideration the fact that although there is no
relative motion
among different points on a rotating disc, each point belongs to a
different noninertial local Fermi frame.
If a rotating ring (or a disc) is constrained to have a
fixed radius from the point of view of an inertial observer, it has been
found that there are tensile stresses in the disc, in agreement with the
prediction of the standard approach. However, contrary to the
prediction of the standard approach, it has been found that an
observer on the rim of the disc will see equal lengths of other differently
moving objects
as an inertial observer whose instantaneous
position and velocity are equal to that of the observer on the rim,
providing that the observations of different events are simultaneous.
This also generalizes to observers arbitrarily moving in
flat spacetime.
The paper deals mainly with flat spacetime, with particular
attention paid to circular motion. However, it gives several
results which are of very general relevance, not only for arbitrary
motion in flat spacetime, but also for general relativity and
curved spacetime.
First, it has been demonstrated that
the generally accepted formula (\ref{eq3}) is not always correct.
The correct definition of the space line element depends on how it
is measured, so (\ref{eq3}) should be used with great care.
In some cases, the ``naive" formula (\ref{dl}) is more appropriate.
One such case is a metric of a frame in flat spacetime that can be
obtained from $g_{\mu\nu}=\eta_{\mu\nu}$ by a transformation of the form of
(\ref{er1})-(\ref{er2}), followed by an arbitrary restricted internal transformation.
Further investigation is needed in order to generalize this result.
Second, the paper demonstrates the importance of the use of
Fermi coordinates. One of the consequences of their use is the result
that for any local observer the velocity of light is isotropic and is
equal to $c$, providing that it is measured by propagating a light
beam in a small neighborhood of the observer.
This fact should be
used for a correct treatment of the Sagnac effect
if one wants to explore the general relativistic corrections.
Fermi coordinates should also be used in order to understand the
physical effects related to a rotating black hole, to give
a correct treatment of the Hawking radiation, as well as
for any other
physical effect, whenever intended to describe the world
how it looks like to a particular observer.
\section*{Acknowledgment}
The author is grateful to Damir Stoi\'{c} for motivating
discussions.
This work was supported by the Ministry of Science and Technology of the
Republic of Croatia under Contract No. 00980102.
|
\section{Introduction}
The unified model for Seyfert galaxies proposes that orientation of a
molecular torus determines the optical emission-line characteristics
(e.g.\ Antonucci\markcite{1} 1993). When the molecular torus lies in
our line of sight, it blocks our view of the broad optical emission
lines, leading to a Seyfert 2 classification. The X-ray emission of
Seyfert 2 galaxies is frequently absorbed (e.g.\ Turner et
al.\markcite{25} 1998; Bassani et al.\markcite{23}
1999), a result which supports this unified model. In the most
extreme case, the obscuring material presents such a high column
density to the observer that no X-rays are transmitted. Such objects
are termed ``Compton-thick'' Seyfert 2s, and the only X-rays observed
from them are ones that have been scattered from surrounding material.
(Diffuse thermal X-rays may also contribute). Such objects are
important because they directly support unified models for Seyfert
galaxies.
The scattering is thought to originate in one or both of two types of
material, each of which imparts characteristic signatures on the
observed X-ray spectrum (for a review, Matt\markcite{15} 1997).
Scattering can occur in the warm optically thin gas that is thought to
produce the polarized broad lines seen in some Seyfert 2s. The
resulting spectrum has the same slope as the intrinsic spectrum with
superimposed emission lines from recombination. This is the process
which appears to dominate in the archetype Seyfert 2 galaxy NGC~1068
(e.g.\ Netzer \& Turner\markcite{20} 1997). Scattering can also occur
in optically thick cool material located on the surface of the
molecular torus. In this case, the process is called Compton
reflection (e.g.\ Lightman \& White\markcite{10} 1988), and the
observed continuum spectrum is flat with superimposed K-shell
fluorescence lines (e.g.\ Reynolds et al.\markcite{22} 1994).
Circinus can be considered the prototype of a Compton-reflection
dominated Seyfert 2 galaxy (Matt et al.\markcite{17} 1996).
Compton-reflection dominated Seyfert 2 galaxies are important because
they may comprise a significant fraction of the X-ray background
(Fabian et al.\markcite{9} 1990). They were once thought to be rare
(e.g. Matt\markcite{15} 1997); however, new observations of objects
selected according to their [\ion{O}{III}] emission-line flux, a
method which ideally does not discriminate against highly absorbed
objects, show that they may be more common than previously thought
(Maiolino et al.\markcite{13} 1998). However, {\it bright} examples
of this class remain rare. This is not surprising, because the
reflected X-rays are very much weaker than the primary continuum.
We present the results of an {\it RXTE} observation of the nearby (18
Mpc) Seyfert 2 galaxy NGC~6300. This object has a flat hard X-ray
spectrum and huge equivalent width iron line which suggests that it is
a Compton-reflection dominated Seyfert 2 galaxy. If so, it is one of
the brightest members of this class known, about half as bright as the
prototype, Circinus, and far brighter than other examples.
\section{Data Analysis}
NGC~6300 was first detected in hard X-rays during a {\it Ginga}
maneuver (Awaki\markcite{2} 1991). We proposed scanning and pointing
observations of this galaxy using {\it RXTE} to confirm the {\it
Ginga} detection. Another Seyfert 2 galaxy, NGC~6393, was detected
during a {\it Ginga} scan and was also investigated as part of this
proposal. The data show that NGC~6393 was very faint ($<0.5\rm
\,counts\, s^{-1}$ in the top-layer for 5 PCUs).
The scanning {\it RXTE} observation of NGC 6300 was performed on 1997
February 14--15. Four of 5 PCUs were on for the entire observation
and analysis was confined to these detectors. A pointed observation
followed on February 20, 1997, performed with all five PCUs on. The
data were reduced using Ftools 4.1 and 4.2 and standard data selection
criteria recommended for faint sources. The resulting exposure for
the pointed observation was 24,896 seconds. NGC 6300 was detected in
all three layers of the PCA, and the top and mid layers were used for
spectral fitting. Background subtraction yielded net count rates for
5 PCUs of 4.4 counts s$^{-1}$ (12.5\% of the total) between 3 and 24
keV for the top layer, and 0.86 counts s$^{-1}$ (8.8\% of the total)
between 9 and 24 keV for the mid layer.
The current standard background model for the {\it RXTE} PCA is quite
good; however, NGC~6300 is a relatively faint source and therefore we
attempt to estimate systematic errors associated with the background
subtraction. Above 30~keV, no signal should be detected; however, we
found a positive signal which could be removed if the background
normalization were increased by 1\%. Below about 7~keV, no signal
should be detected in the mid or bottom layers. We observed a small
deficit in signal which would be removed if the background
normalization were decreased by 1\%. We consider this evidence that
the systematic error on the background subtraction is less than 1\%.
The scan observation consisted of 4 passes over the object with a
total scan length of 6 degrees. The scans were performed keeping the
declination constant during the first two passes and the right
ascension constant during the second two passes. The resulting scan
profiles when compared with the optical position clearly
indicate that NGC~6300 is the X-ray source. The field of view of the
PCA is less than two degrees in total width. Since the scan length
was 6 degrees, and since there are apparently no other X-ray sources
in the field of view, the ends of the scan paths can be used to check
the quality of the background subtraction. This was of some concern
since NGC~6300 is rather near the Galactic plane ($l=328$, $b=-14$)
and thus there could be Galactic X-ray emission not accounted for in
the background model. We accumulated spectra with offset from the
source position $>1.5^\circ$. The exposure time was 1472 seconds.
The count rate between 3 and 24 keV was $-0.28
\pm 0.19 \rm\, counts\,s^{-1}$, so there was no evidence for
unmodeled Galactic emission. Thermal model residuals show
no pattern; i.e., there is no evidence for a 6.7~keV iron emission
line from the Galactic Ridge (Yamauchi \& Koyama\markcite{28} 1993).
\begin{figure}[t]
\vbox to3.0in{\rule{0pt}{3.0in}}
\special{psfile=ngc6300_fig1.ps angle=270 hoffset=10 voffset=250 vscale=100
hscale=100}
\caption{Ratio of data to a model consisting of power law plus
Galactic absorption model. The excess near 6.4~keV clearly indicates
the large equivalent width iron line. Solid points and crosses denote
the top-layer and mid-layer data, respectively.}
\end{figure}
The spectrum from the pointed observation was first modeled using a
power law plus Galactic absorption set equal to $9.38\times
10^{20}\rm\, cm^{-2}$ (Figure 1; Dickey \& Lockman\markcite{6} 1990).
This model did not fit the data well ($\chi^2=609$ for 82 degrees of
freedom (d.o.f.)). The photon index is very flat ($\Gamma=0.60$),
there is clear evidence for an iron emission line and there are
negative low-energy residuals. Addition of a narrow ($\sigma=0.05\rm
\,keV$) line with energy fixed at $6.4\rm\, keV$ improves the fit
substantially ($\Delta\chi^2=473$) but low energy residuals remain.
Additional absorption in the galaxy rest frame improves the fit
substantially ($\Delta\chi^2=45$). Freeing the line energy again
improves the fit ($\Delta\chi^2=11$); the best fit rest-frame line
energy is $6.26 \rm\,keV$. Freeing the line width marginally improves
the fit ($\Delta\chi^2=5$). The final fit parameters are listed in
Table 1 and fit results are shown in Figure 2.
The absorbed power law plus iron line is an acceptable model. The
notable properties of the fit are a very flat photon index
($\Gamma=0.68$) and very large equivalent width ($920\,\rm eV$). Such
parameters suggest that the spectrum of NGC~6300 is dominated by
Compton-reflection (Matt et al.\markcite{17} 1996; Malaguti et
al.\markcite{12} 1998; Reynolds et al.\markcite{22} 1994). We next use
the {\it pexrav} model in XSPEC to explore this possibility. This
model calculates the expected X-ray spectrum when a point source of
X-rays is incident on optically thick, predominately neutral (except
hydrogen and helium) material. The parameter $R$ measures the solid
angle $\Omega$ subtended by the optically thick material:
$R=\Omega/2\pi$. The model that was fit includes a narrow iron line
and a direct and reflected power law; additional absorption also
appears to be necessary. The low resolution and limited band pass
provided by the {\it RXTE} spectrum means that not all of the model
parameters could be constrained by the data; thus, the energy of the
exponential cutoff was fixed at 500~keV, approximately the value that
has been found in {\it OSSE} data from Seyfert galaxies (Zdziarski et
al.\markcite{29} 1995), and the inclination was initially fixed
arbitrarily at $45^\circ$. This model provided a fairly good fit to
the data ($\chi^2=112$ for 77 d.o.f.). However, the best fit value of
$R$ is very large (1450) and not well constrained. This indicates
that the spectrum can be modeled using reflection alone, and there is
no significant contribution of direct emission (e.g.\ Matt et al.\markcite{17}
1996).
Reflection alone gives the same $\chi^2$ as the model which includes a
weak direct component, but the fit is still not completely
satisfactory, and it could be improved in two ways. The first is to
allow the iron abundance to vary. We set the iron abundance relative
to solar in the {\it pexrav} model equal to the abundances of light
elements and allow these parameters to vary together. The fit is
improved significantly ($\Delta\chi^2=-36$ for $\Delta$d.o.f.=1) and
gives a slightly subsolar abundance. The second is to allow the
inclination to be free. The fit is not very sensitive to this
parameter, as $\Delta\chi^2$ over the whole range is 9.2. The best
fit value is $\cos(\Theta)=0.22$, corresponding to 77$^\circ$ from the
normal. The parameters are listed in Table 1.
\begin{deluxetable}{ll}
\small
\tablewidth{20pc}
\tablenum{1}
\tablecaption{Spectral Fitting Results}
\tablehead{
\colhead{Parameter} & \colhead{Value} \\}
\startdata
\multicolumn{2}{c}{Power Law Model} \nl
$\Gamma$ & $0.68_{-0.09,-0.13}^{+0.09,+0.16}$ \nl
$\rm N_H$ ($10^{22}\rm \,cm^{-2}$) &
$5.2_{-1.7,-2.7}^{+1.8,+3.5}$ \nl
$\rm E_{Fe}$ (keV) & $6.26_{-0.06,-0.02}^{+0.06,+0.02}$ \nl
$\rm \sigma_{Fe}$ (keV) & $0.32_{-0.16,+0.07}^{+0.13,-0.07}$ \nl
$\rm F_{Fe}$ ($10^{-5}\,\rm \,cm^{-2}\,s^{-1}$) &
$8.2_{-1.1,+0.5}^{+1.2,-0.4}$ \nl
$\rm EW_{Fe}$ (eV) & $920_{-130,+100}^{+140,-90}$ \nl
$\chi^2$/79 d.o.f. & $75.5_{\ldots,-2.2}^{\ldots,+6.7}$ \nl
\tableline
\multicolumn{2}{c}{Compton Reflection Model} \nl
$\Gamma$ & $1.89_{-0.09,+0.08}^{+0.08,-0.13}$ \nl
Abundance $^a$ & $0.61_{-0.11,+0.15}^{+0.11,-0.16}$ \nl
$\cos(\Theta)$ & $0.22_{-0.22,+0.02}^{+0.16,-0.04}$ \nl
$\rm E_{Fe}$ (keV) & $6.29_{-0.08,+0.02}^{+0.09,-0.02}$ \nl
$\rm \sigma_{Fe}$ (keV) & $0.22_{-0.22,-0.13}^{+0.20,+0.06}$ \nl
$\rm F_{Fe}$ ($10^{-5}\,\rm \,cm^{-2}\,s^{-1}$) &
$4.7_{-1.0,-1.2}^{+1.2,+1.2}$ \nl
$\rm EW_{Fe}$ (eV) & $470_{-100,-150}^{+120,+200}$ \nl
$\chi^2$/77 d.o.f. & $70.3_{\ldots,+9.2}^{\ldots,-1.6}$ \nl
\tableline
\multicolumn{2}{c}{Dual Absorber Model} \nl
$\Gamma$ & $1.71^{+0.21,+0.27}_{-0.19,-0.21}$ \nl
$\rm N_{H(thin)}$ ($10^{22}\rm\,cm^{-2}$) &
$7.7_{-5.2,-1.4}^{+4.0,+2.0}$ \nl
$\rm N_{H(thick)}$ ($10^{22}\rm\,cm^{-2}$) &
$58_{-22,-2}^{+23,+5}$ \nl
$\rm A_{thick}/A_{thin}$$^b$ & $1.9_{-0.9,-0.7}^{+1.4,+1.0}$ \nl
$\rm E_{Fe}$ (keV) & $6.26_{-0.12,+0.002}^{+0.09,-0.005}$ \nl
$\rm \sigma_{Fe}$ (keV) & $0.23_{-0.23,+0.04}^{+0.25,-0.05}$ \nl
$\rm F_{Fe}$ ($10^{-5}\,\rm \,cm^{-2}\,s^{-1}$) &
$3.2_{-1.5,+0.8}^{+1.4,-0.7}$ \nl
$\rm EW_{Fe}$ (eV) & $470_{-220,+50}^{+210,-60}$ \nl
$\chi^2$/79 d.o.f. & $67.6_{\ldots,-1.2}^{\ldots,+2.4}$ \nl
\enddata
\tablecomments{Two kinds of errors are given for each parameter value. The
first one is the statistical error which represents 90\% confidence
for one parameter of interest ($\Delta\chi^2=2.71$). The second one
is an estimate of the systematic error obtained by changing the
normalization of the background. The results of background
under- and oversubtraction by 1\% are given in the sub- and
superscript, respectively.}
\tablenotetext{a}{Fraction of solar abundance in iron and light
elements.}
\tablenotetext{b}{Ratio of power-law normalizations.}
\end{deluxetable}
We investigated the choice of fixed parameters in the {\it pexrav}
model {\it a posteriori}. Increasing the cutoff energy did not change
the fit. Decreasing the cutoff to 100~keV produced small differences
in the parameters; namely, the photon index was smaller, the abundance
was higher, and the line flux was lower, but the differences are
within the statistical errors of the adopted model. We also
investigated the situation when the iron abundance was allowed to
vary, but the abundances of lighter metals were maintained at the
solar value. A significantly larger photon index by $\Delta\Gamma
\approx 0.15$ was required, due to the decreased reflectivity in soft
X-rays. We also investigated the effect of a 1\% systematic error in
the background normalization, and the results are listed in Table 1.
This resulted in the largest change in the fit parameters, but the
resulting estimated systematic errors are in the worst case less than
a factor of two larger than the statistical errors.
Because of the low energy resolution of the {\it RXTE} spectra, other
models can be found which fit equally well. It is possible to describe
the spectra using a sum of absorbed power laws (the ``dual absorber''
model; e.g.\ Weaver et al.\markcite{27} 1994). Specifically, the
model consisted of two power laws, both absorbed by a moderate column,
and one absorbed by a heavy column. The resulting photon index was
very flat ($\Gamma=1.15$), and is therefore deemed unphysical.
However, including reflection with $R=1$ in the dual absorber model
gave an insignificant improvement in fit ($\Delta\chi^2=-0.6$) but a
more plausible photon index ($\Gamma=1.7$). The fit parameters are
given in Table 1. It is notable that the spectra cannot be described
using a highly absorbed transmitted component and an unabsorbed
Compton-reflection dominated component, as has been found to be
appropriate for Mrk~3 (Cappi et al.\markcite{4} 1999).
\begin{figure}[t]
\vbox to4.0in{\rule{0pt}{4.0in}}
\special{psfile=ngc6300_fig2.ps angle=270 hoffset=-40 voffset=420 vscale=70
hscale=70}
\caption{Unfolded spectra and ratio of data to model for the three
models considered (see Table 1).}
\end{figure}
\section{Discussion}
\subsection{Compton Reflection Continuum Model}
The X-ray continuum of NGC~6300 can be modeled as pure Compton
reflection. For solar abundance, the iron line equivalent width
relative to the reflection continuum is predicted to be between 1 and
2 keV depending on inclination (Matt, Perola \& Piro\markcite{18}
1991). When we fit a power law continuum to the spectra, the
observed equivalent width is nearly 1~keV. However, when the
reflection continuum is fitted, the measured equivalent width is
reduced to 470~eV. The reason for the reduction in the measured
equivalent width is that the reflection continuum model includes a
substantial iron edge. In low resolution data, the iron
line and iron edge overlap in the response-convolved spectra.
Therefore, when the continuum is modeled by a power law, the iron line
models both the line and the edge, so the measured equivalent width is
larger than when the continuum is modeled by the reflection continuum
which includes the iron edge explicitly. This effect can be seen in
Figure 2.
A significant improvement is obtained when the iron abundance in the
reflection continuum model is allowed to be subsolar. The iron
abundance is determined by the depth of the iron edge. Therefore, the
subsolar abundance fits because the iron edge is apparently not as
deep as the model predicts. The fact that the iron line has a lower
equivalent width than predicted in the Compton reflection continuum
model may also support subsolar abundance. Alternatively, however,
the apparent subsolar abundances may at least partially be due to
calibration uncertainties in the {\it RXTE} PCA. (The resolution of
the {\it RXTE} PCA is under some debate; see, Weaver, Krolik \&
Pier\markcite{26} 1998). The iron line and edge overlap in
the response-convolved spectra. If the true energy resolution is worse
than the current estimated value, then because the line is an excess
and the edge is a deficit, both would be measured to be smaller than
they really are.
Another source of uncertainty may come from the models themselves,
which depend strongly on the geometry of the illuminating and
reprocessing material. Models generally assume a point source of
X-rays located above a disk and illuminating it with high covering
fraction. Such an ideal case may not be attained in nature.
\subsection{Dual Absorber Continuum Model}
The dual absorber model can also describe the spectra. That
such a fit is successful is not surprising, as any flat continuum can
be modeled as a sum of absorbed power laws (e.g.\ the X-ray
background).
The iron line equivalent width for the dual absorber model is similar
to that found for the Compton-reflection model, and smaller than that
found for the power law model. The reason for the difference is that,
like the Compton reflection continuum model, the dual absorber model
explicitly includes an iron edge. A plausible origin for the iron
line in the dual absorber model is in the absorbing material itself.
However, the iron line equivalent width appears to be too large to
have been produced in the absorbing material. We investigate this
possibility by comparing the observed iron line flux to the predicted
value from a spherical shell of gas surrounding an isotropically
illuminating point source (Leahy \& Creighton\markcite{11} 1993). A
line flux of $2.3\times 10^{-5}\rm \, photons\,cm^{-2}\,s^{-1}$ is
predicted for the absorption columns and covering fractions determined
by the fit; this is about half of what is observed ($4.7\times
10^{-5}\rm \, photons\,cm^{-2}\,s^{-1}$). The dual absorber model
also requires a reflection component with $R=1$ and therefore an
additional iron line with equivalent width $\sim 100$~eV is expected
from the reflection. Then the predicted flux increases to $3.3\times
10^{-5}\rm \, photons\,cm^{-2}\,s^{-1}$, about 70\% of what is
observed. The predicted line flux would be smaller if the absorbing
material does not completely cover the source, a circumstance that
would exacerbate the difference between predicted and observed flux.
Thus, the iron line equivalent width, at least to first
approximation, appears to be too large to have been produced in the
absorbing material required by the dual absorber model. Therefore,
either an iron overabundance is required in the dual absorber model,
or the alternative model, the Compton-reflection dominated model, is
favored.
Discrimination between models will come with observations using
detectors with better energy resolution. If NGC~6300 is a
Compton-thick Seyfert 2, then we should detect soft X-ray emission
lines (e.g.\ Reynolds et al.\markcite{22} 1994). The lack of any
observable soft excess in the {\it RXTE} data may imply that there is
absorption by the host galaxy (see below), or that there is little
contamination from thermal X-rays or scattering from warm gas. If the
latter case is true, NGC~6300 will be a particularly clean example of
a Compton reflection-dominated Seyfert 2 galaxy, and the observed soft
X-ray emission lines should be unambiguously attributable to
fluorescence. Such proof should be easily attainable as NGC~6300 is
bright compared with known Compton-reflection dominated Seyfert 2
galaxies.
\subsection{Information from Other Wavebands}
Optical observations provide some support for Compton-thick absorption
in NGC 6300. Intrinsically, both hard X-rays and forbidden optical
emission lines should be emitted approximately isotropically;
therefore, the ratio of these quantities should be the same from
object to object. However, if the absorption is Compton-thick, the
observed hard X-ray luminosity and therefore $L_X/L_{[OIII]}$ will be
significantly reduced; the ratio of the power law to reflection
component 2--10 keV fluxes in the {\it pexrav} model is $\approx 15$.
Care must be taken when applying this test, as the [\ion{O}{III}] flux
must be corrected for reddening in the narrow-line region, and
determination of the reddening using narrow-line Balmer decrements can
be difficult. The optical spectrum of NGC~6300 is dominated by
starlight (e.g.\ Storchi-Bergmann \& Pastoriza\markcite{24} 1989); to
remove the Balmer absorption from the stars, an accurate galaxy
spectrum subtraction must be done. Another complication could be
narrow Balmer lines from star formation. NGC~6300 has a well-studied
starburst ring (e.g.\ Buta\markcite{3} 1987), but H$\alpha$ images
show that the \ion{H}{II} regions are located $>0.5^\prime$
from the nucleus, with a little diffuse emission inside
of that (e.g.\ Evans et al.\markcite{7} 1996). High quality long-slit
spectra yield $A_v \approx 2.5-3$ from both the red continuum and the
Balmer decrement (Storchi-Bergmann 1999, P.\ comm.). The observed
2--10 keV flux is $6.4\times 10^{-12}\rm
\,erg\,cm^{-2}\,s^{-1}$ (corresponding to a luminosity of $2.5\times
10^{41}\,\rm erg\,s^{-1}$). Then $L_X/L_{[OIII]}$ is approximately
1.1--1.9. This value is quite low and comparable to those obtained from
Compton-thick Seyfert 2s by Maiolino et al.\markcite{13} (1998). In
particular, Circinus has $F_X=1.4\times 10^{-11} \rm
ergs\,cm^{-2}\,s^{-1}$ (Matt et al.\markcite{16} 1999) and
reddening-corrected $F_{[OIII]}=1.95\times 10^{-11}$ ($A_v=5.2 \pm
0.4$; Oliva et al.\markcite{21} 1994), yielding $L_X/L_{[OIII]}=0.7$.
For comparison, a reddening-corrected sample of Seyfert 1s taken from
Mulchaey et al.\markcite{19} 1994 as a mean ratio of 14.8 ($1\sigma$ range
6.4--33.9). In contrast, the intrinsic luminosity (i.e.\ corrected
for absorption) for the dual absorber model is $6.9\times 10^{41}\,\rm
erg\,s^{-1}$, implying $L_X/L_{[OIII]} \approx $3.0--5.2.
It is notable also that NGC~6300 shows the reddest continuum toward
the nucleus in a sample of objects studied with long slit spectroscopy
(Cid Fernandes, Storchi-Bergmann \& Schmitt\markcite{5} 1998).
Furthermore, there seems to be a correlation between the presence of a
bar in the host galaxy and presence of a Compton thick Seyfert 2
nucleus (Maiolino, Risaliti \& Salvati\markcite{14} 1999); NGC 6300
has a bar (e.g.\ Buta\markcite{3} 1987).
\subsection{Consistency with Einstein IPC Observation}
The X-ray spectra from Seyfert 2 galaxies frequently includes a soft
spectral component which may originate in scattering by warm gas or
diffuse thermal X-rays. However, lack of detection in an {\it
Einstein} IPC observation, combined with the {\it RXTE} observation
presented here, shows that there is no evidence for such a component
in NGC~6300.
In 1979 NGC~6300 was observed with the {\it Einstein} IPC for 990
seconds. It was not detected and the three sigma upper limit to the
count rate was $1.19\times 10^{-2}\rm\, counts\,s^{-1}$ between 0.2
and 4.0 keV (Fabbiano, Kim, \& Trinchieri\markcite{8} 1992). The
Compton reflection-dominated model predicts a count rate of $1.3\times
10^{-2}\rm\, counts\,s^{-1}$, just the same order as the upper limit,
and probably consistent within the uncertainties of the model and the
relative flux calibrations of the two instruments.
There may be intrinsic absorption in the system, for example, from the
host galaxy. The observed optical reddening of $A_v=2.5-3.0$
corresponds to an absorption column of 4--5$\times 10^{21}\rm
cm^{-2}$, assuming a standard dust to gas ratio. Including this
column in the Compton-reflection dominated model leads to a predicted
IPC flux of $1.1\times 10^{-2}\rm\, counts\,s^{-1}$, consistent with
the upper limit. Because the {\it RXTE} band pass is truncated at
3~keV, it is impossible to estimate with accuracy the intrinsic
absorption. For the Compton-reflection model, the 90\% upper limit
for one parameter of interest is $1.8\times 10^{22}\rm\,cm^{-2}$.
Alternatively, there may have been variability within the 18
years between the IPC and the {\it RXTE} observation. A probable site
for the reflection is the inner wall of the molecular torus
which blocks the line of sight to the nucleus. In unified models for
Seyfert galaxies, this material is located outside of the broad line
region, and can be 1--100 pc from the nucleus. Short term variability
is not expected; long term variability is possible but requires a long
term trend in flux.
\acknowledgements
KML acknowledges useful discussions on the {\it RXTE} background with
Keith Jahoda. KML gratefully acknowledges support by NAG-4112 ({\it
RXTE}) and NAG5-7971 (LTSA).
|
\section{Introduction}
Strangeness as a possible signature of the phase transition from a
hadronic state to a QGP state was put forward about 16 years ago
\cite{rafelski1}. It was based on the prediction that the
production of strange quark pairs would be enhanced as a result of
the approximate chiral symmetry restoration in a QGP state in
comparison with a hadronic state. The strangeness enhancement in
pA and AA collisions with respect to the superposition of
nucleon-nucleon collisions has been investigated and confirmed by
many experimental groups \cite{wa85,na351,na36,wa97,naga}.
However, alternative explanations exist, they are based on the `conventional'
physics in the hadronic regime, like rescattering, string-string
interaction, etc. \cite{satai,venus,rqmd}. The first
detailed theoretical study of strangeness production can be found
in \cite{rafelski2}, where the enhanced relative yield of strange
and multi-strange particles in nucleus-nucleus collisions with
respect to proton-nucleus interactions has been suggested as a
sensitive signature of a QGP.
We have done a series of studies in recent years investigating
strangeness enhancement with a hadron and string scenario
\cite{satai,satai2,satai3,taisa1,taisa2}, from which a Monte-Carlo
event generator, LUCIAE, was developed \cite{luciae}. Those
studies indicate that including rescattering of the final state
hadrons is still not enough to reproduce the NA35 \cite{na351} data of
strange particle production. To reproduce the NA35 data needs to
rely further on the mechanism of reduction of the strange quark
suppression in string fragmentation, which contributes to the enhancement of
strange particle yield in nucleus-nucleus collisions with respect
to the superposition of the nucleon-nucleon collisions
\cite{satai2,satai3,taisa1,taisa2}. Similarly, in order to
reproduce the NA35 data, the RQMD generator, equipped with
rescattering though, has to resort to the colour rope mechanism
\cite{rqmd}. In this picture it is assumed that the neighboring
interacting strings might form a string cluster called colour rope
in pA and AA collisions. The colour rope then fragments in a
collective way and tends to enhance the production of the strange
quark pairs from the colour field of strings through the increase
of the effective string tension.
It has been known for years that the strange quark suppression
factor ($\lambda$ hereafter), i.e., the suppression of s quark
pair production in the color field with respect to u or d pair
production, in hadron-hadron collisions is not a constant, but
energy-dependent, increasing from a value of 0.2 at the ISR
energies to about 0.3 at the top of the SPS energies \cite{kapa}.
In \cite{taisa1} we proposed a mechanism to investigate the energy
dependence of $\lambda$ in hh collisions by relating the effective
string tension to the production of hard gluon jets (mini-jets). A
parameterization form was then obtained, which reproduces the
energy dependence of $\lambda$ in hh collisions reasonably well
\cite{taisa1}. When the same mechanism is used in the study of
pA and AA collisions it is found that $\lambda$ would increase
with the increase of energy, mass and centrality of a colliding system
as a result of mini-jet(gluon) production stemming from the
string-string interaction. Our model reproduced nicely the data of
strange particle production in hh \cite{taisa1}, pA, and AA
\cite{satai3,taisa2} collisions.
In this work we use above ideas \cite{taisa1,taisa2} to study the
recently published WA97 data of the enhanced production of singly
and multiply strange particles in p-Pb and Pb-pb collisions at
158A GeV/c. The study indicates that the WA97 data, which
revealed that the enhancement of strange particle yield
increases with the increasing of centrality and of s quark content
in multiply strange particles in Pb-Pb collisions with respect
to p-Pb collisions, could be explained in a hadron-string model
except for $\Omega$ yield in the Pb-Pb data.
\section{Brief review of the LUCIAE model}
LUCIAE model is developed based on the FRITIOF model
\cite{fritiof}. FRITIOF is a string model, which started from the
modeling of inelastic hadron-hadron collisions and it has been
successful in describing many experimental data from the low
energies at the ISR-regime all the way to the SPS energies
\cite{B.N,H.P1}. In this model a hadron is assumed to behave like
a massless relativistic string. A hadron-hadron collision is
pictured as the multi-scattering of the partons inside the two
colliding hadrons. In FRITIOF, during the collision two hadrons
are excited due to longitudinal momentum transfers and/or a
Rutherford Parton Scattering (RPS). The highly excited states will
emit bremsstrahlung gluons according to the soft radiation model.
They are afterwards treated as excitations i.e. the Lund Strings
and allowed to decay into final state hadrons according to the
Lund fragmentation scheme.
The FRITIOF model has been extended to also describe
hadron-nucleus and nucleus-nucleus collisions by assuming that the
reactions are superposition of binary hadron-hadron collisions in
which the geometry of the nucleus plays an important role because
the nuclei should then behave as a ``frozen'' bag of nucleons.
However in the relativistic nucleus-nucleus collisions there are
generally many excited strings formed close by each other during a
collision. Thus in the LUCIAE model a Firecracker model
\cite{fire} is proposed to deal with the string-string collective
interaction. In the Firecracker model it is assumed that several
string from a relativistic heavy ion reaction will form a cluster
and then the strings inside such a cluster will interact in a
collective way. We assume that the groups of neighbouring strings
in a cluster may form interacting quantum states so that both the
emission of gluonic bremsstrahlung as well as the fragmentation
properties can be affected by the large common energy density.
In relativistic nucleus-nucleus collision there are generally a
lot of hadrons produced, however, FRITIOF does not include the
final state interactions. Thus in LUCIAE a rescattering model
\cite{satai} is devised to consider the reinteraction of the
produced hadrons with each other and with the surrounding cold
spectator matter. The distributions of the final state hadrons
will be affected by the rescattering process. We refer to the
Refs. \cite{satai,luciae} for the details and we just give here
the list of the reactions involving in LUCIAE, which are cataloged
into
\begin{tabbing}
ttttttttttttttt\=ttttttttttttttt\=tttttt\=tttttttttttttttt\= \kill
\>$\pi$$N$$\rightleftharpoons$ $\Delta$$\pi$
\> \>$\pi$$N$$\rightleftharpoons$ $\rho$$N$\\
\> $N$$N$$\rightleftharpoons$ $\Delta$$N$
\> \>$\pi\pi \rightleftharpoons k\bar{k}$\\
\>$\pi N \rightleftharpoons kY$
\> \>$\pi\bar{N} \rightleftharpoons \bar{k}\bar{Y}$\\
\>$\pi Y \rightleftharpoons k\Xi$
\> \>$\pi\bar{Y} \rightleftharpoons \bar{k}\bar{\Xi}$\\
\>$\bar{k}N \rightleftharpoons \pi Y$
\> \>$k\bar{N} \rightleftharpoons \pi\bar{Y}$\\
\>$\bar{k}Y \rightleftharpoons \pi\Xi$
\> \>$k\bar{Y} \rightleftharpoons \pi\bar{\Xi}$\\
\>$\bar{k}N \rightleftharpoons k\Xi$
\> \>$k\bar{N} \rightleftharpoons \bar{k}\bar{\Xi}$\\
\>$\pi\Xi \rightleftharpoons k\Omega^- $
\> \>$\pi\bar{\Xi} \rightleftharpoons \bar{k}\overline{\Omega^-}$\\
\>$k\bar{\Xi} \rightleftharpoons \pi\overline{\Omega^-}$
\> \>$\bar{k}\Xi \rightleftharpoons \pi\Omega^-$\\
\>$\bar{N}N$ annihilation\\
\>$\bar{Y}N$ annihilation\\
\end{tabbing}
where $Y$ refers to the $\Lambda$ or $\Sigma$ and $\Xi$ refers to the
$\Xi^-$ or $\Xi^0$. There are 364 reactions involved altogether.
In addition, the reduction mechanism of s quark suppression, i. e., the s
quark suppression factor increasing with energy, centrality, and mass of the
colliding system, which is linked to string tension, is included in
LUCIAE via the parameterized formulas \cite{taisa1,taisa2}
\begin{equation}
\kappa_{eff}=\kappa_{0} (1-\xi)^{-\alpha},
\label{f2}
\end{equation}
where $\kappa_{0}$ is the string tension of the pure $q\bar{q}$ string,
$\alpha$ is a parameter $\sim$ 3, and $\xi$ ($\leq$ 1) is calculated by
\begin{equation}
\xi =\frac{\ln(\frac{k_{\perp max}^2}{s_{0}})}{\ln (\frac{s}{s_{0}}) +
\sum_{j=2}^{n-1} \ln (\frac{k_{\perp j}^2}{s_{0}})},
\label{f3}
\end{equation}
which represents the scale that a multigluon string is deviated from a pure
$q\bar{q}$ string.
The s quark suppression factor, $\lambda$, of two string states can thus
be calculated by
\begin{equation}
\lambda_{2} = \lambda_{1}^ {\frac{\kappa_{eff1}}{\kappa_{eff2}}},
\label{f1}
\end{equation}
where $\kappa_{eff}$ refers to the effective string tension of
a multigluon string. Since $\lambda$ is always less than one, above
equation indicates the larger effective string tension the more
reduction of s quark suppression. The effective string tension is
then relevant to the hard gluon kinks (mini-(gluon) jets) created
on the string.
It should be mentioned that the LUCIAE (FRITIOF) event generator
runs together with JETSET routine. In JETSET routine there are
model parameters PARJ(2) (i.e., $\lambda$) and PARJ(3). PARJ(3) is
the extra suppression of strange diquark production compared to
the normal suppression of strange quark pair. Both PARJ(2) and
PARJ(3) are responsible for the s quark (diquark) suppression and
related to the effective string tension (the relation of Eq. (3)
holds true for PARJ(3) as for $\lambda$). Besides $\lambda$ and PARJ(3) there
is PARJ(1), which stands for the suppression of diquark-antidiquark
pair production in the color field in comparison with the
quark-antiquark pair production and is related to the effective
string tension as well. The mechanism mentioned above is performed
via these parameters in program. How these three parameters affect
the multiplicity distribution of final state particles can be
found in \cite{satai2,satai3}.
\section{Results and discussions}
In Table 1 is given the results of the JETSET parameters PARJ(1),
PARJ(2) (i.e., $\lambda$), and PARJ(3) varying with the centrality
and the size of collision system in p-Pb and Pb-Pb collisions
at 158A GeV/c. That seems quite reasonable.
Fig. 1a shows the calculated $\Lambda+\bar{\Lambda}$, $\Xi^-+
\overline{\Xi^-}$, and $\Omega^-+\overline{\Omega^-}$ yields per event
($|y-y_{cm}| \leq$ 0.5 and p$_T$ $\geq$ 0 GeV/c) as a function of the
number of participant in minimum bias p-Pb collisions and in central
(b=2) Pb-Pb collisions at 158A GeV/c (open labels) comparing
with WA97 data (full labels) \cite{wa97}. The corresponding
results in Pb-Pb collisions after recaling each yield according
to its value in p-Pb are given in Fig. 1b. One knows from Fig. 1a that
the agreement between theory and experiment is quite well
for $\Lambda+\bar{\Lambda}$ and $\Xi^-+\overline{\Xi^-}$, however,
for $\Omega^-+\overline{\Omega^-}$ the theoretical results are lower
than experiments. That should be study further both theoretically
and experimentally. In Fig. 1b the theoretical results of
$\Omega^-+\overline{\Omega^-}$ are also lower than experiments,
however, the trend of the strangeness enhancement increasing with
increase of the centrality and of the s quark content in strange
particles is reproduced quite well.
In Fig. 2 and 3 are given, respectively, the calculated m$_T$
spectra ($|y-y_ {cm}| \leq$ 0.5) of $\Lambda$, $\bar{\Lambda}$,
$\Xi^-$, $\overline{\Xi^-}$ and $\Omega^-+\overline{\Omega^-}$ in
p-Pb and Pb-Pb collisions at 158A GeV/c (open labels). The
corresponding full labels in those figures are the corresponding WA97
data \cite{wa97}. One sees from figure 2 that the agreement between
theory and experiment is reasonably good, except that the fluctuation
in theoretical results of $\Omega^-+\overline{\Omega^-}$ m$_T$
spectrum has to be improved. However, the situations in figure 3
is much better, i.e., the agreement between theory and experiment
is reasonably good.
In summary, we have used a hadron and string cascade model,
LUCIAE, to investigate the WA97 data of the strangeness
enhancement increasing with the increase of the centrality and of
the s quark content in strange particles. Relying on the mechanism of
the reduction of s quark suppression in string fragmentation leads
to the enhancement of strange particle yield in nucleus-nucleus
collisions the WA97 data could be reproduced nicely except $\Omega$ yield
in Pb+Pb collisions, which need to be studied further.
\section{ACKNOWLEDGMENTS}
We would like to thank T. Sj\"{o}strand for detailed instructions
of using PYTHIA. This work was supported by national Natural
Science Foundation of China and Nuclear Industry Foundation of
China.
|
\subsection{Two Painlev\'e\ branches and the singular manifold method}
These equations (\ref{1.2},\ref{1.3}) have the property that the leading
analysis of the expansion
(\ref{1.1}) provides $\alpha=1$ and $a_0=\pm 1$, where the $\pm$ sign in
$a_0$ indicates that
there are two possible Painlev\'e\ branches.
Equations with two Painlev\'e\ branches and how to extend the SMM to them have been
studied in previous papers (cf.\ \cite{EG93,EG97}). According to these,
when an equation,
such as (\ref{1.2}) or (\ref{1.3}), has two Painlev\'e\ branches, then the
truncation of (\ref{1.1})
should be made for both branches simultaneously. This means that we should
work with truncated
solutions $\~y$ of the following form \cite{EG97}
\begin{equation} \~y=y+{g'\over g}-{h'\over h},\label{1.4}\end{equation}
where $y(x)$ is also a solution of the equation, and $g(x) =0$ is the
singularity for the $+$
expansion and $h(x)=0$ the singularity for the $-$ expansion. Henceforth we
shall call $g$ and
$h$ {\it singular manifolds} \cite{Weiss83}.
The SMM usually requires that each coefficient in the different powers of
the singular manifold
that arises from substitution of the truncated expansion in the
differential equation be equal to
zero. Due to the non-linearity of PII and PIV, the substitution of
(\ref{1.4}) in (\ref{1.2}) and
(\ref{1.3}) respectively provides terms that mix the powers in $h$ and $g$.
To solve this problem
we use a {\it decoupling ansatz} \cite{EG97} given by
\begin{equation} {g'\over g}{h'\over h}=A{g'\over g}+B{h'\over
h},\label{1.5}\end{equation}
where $A$ and $B$ are functions of $y$, and $g$ and $h$ to be determined
from the truncation
itself. Furthermore, taking the derivative of (\ref{1.5}) with respect to
$x$ we have
\begin{equation}
A'=A(r-A-B)\qquad \qquad B'=B(v-A-B),\label{1.6}\end{equation}
where $v={g''/g'}$ and $r={h''/h'}$.
Within this framework, the objective of this paper is to prove that for PII
and PIV the improved
version of the SMM including two singular manifolds provides the following
results:
\begin{itemize}
\item[(i)] Modified versions of PII and PIV (namely mPII and mPIV).
\item[(ii)] Miura transformations between PII and mPII and between PIV and
mPIV.
\item[(iii)] B\"acklund\ transformations for mPII and mPIV.
\item[(iv)] By regarding the parameter $\ell$ as a discrete variable,
discrete equations
can be derived associated with mPII and mPIV.
\item[(v)] Solutions of these discrete equations allow us to construct
solutions of PII and
PIV as a ``linear" superposition of solutions of the discrete equations.
\end{itemize}
\setcounter{equation}{0{Painlev\'e\ II}
Substituting (\ref{1.4}) into PII (\ref{1.2}) and the use of (\ref{1.5}) to
decouple the
crossed terms yields
\begin{eqnarray}
A&=&y+\tfr{1}{2}v,\label{2.2}\\
B&=&-y+\tfr{1}{2}r,\label{2.3}\\
0&=&v'+v^2-6y^2-x+6A(2y+B-A),\label{2.4}\\
0&=&r'+r^2-6y^2-x-6B(2y+B-A).\label{2.5}\end{eqnarray}
Using (\ref{1.6}) in (\ref{2.2}--\ref{2.3}) gives
$$AB=k, $$
where $k$ is a constant. By substituting (\ref{2.2}--\ref{2.3}) into
(\ref{2.4}--\ref{2.5}), the
following singular manifold equations are obtained
\begin{equation} v'-\tfr12{v^2}+6k-x=0,\qquad\qquad
r'-\tfr12{r^2}+6k-x=0.\label{2.7}\end{equation}
\subsection{A Miura transformation for modified PII}
If we return to equation (\ref{1.6}), substituting $A$ and $B$ as given by
(\ref{2.2}) and
(\ref{2.3}), respectively, and using (\ref{2.7}), we have
\begin{equation}
y'-y^2-\tfr12{x}+v'+2k=0,\qquad-y'-y^2-\tfr12{x}+r'+2k=0,\label{2.8}\end{equation}
which means that the expression $y'-y^2-\tfr12{x}$ depends only on the
singular manifold $g$
whilst $-y'-y^2-\tfr12{x}$ depends only on $h$. Hence it is useful to
define the following
functions $m'$ and $n'$ as follows
\begin{equation} 2m'=y'-y^2-\tfr12{x},\qquad
2n'=-y'-y^2-\tfr12{x}.\label{2.9}\end{equation}
With the aid of (\ref{1.2}), these equations (\ref{2.9}) can be integrated
to give
\begin{equation} 2m=(y')^2-(y^2+\tfr12{x})^2-(2\ell-1)y,\qquad
2n=(y')^2-(y^2+\tfr12{x})^2-(2\ell+1)y,\label{2.10}\end{equation}
respectively. In order to identify the equations that $m$ and $n$ satisfy,
we take the derivative
of (\ref{2.9}), which gives
\begin{equation} 2m''=-4ym'+\left(\ell-{\tfr12}\right),\qquad
2n''=4yn'-\left(\ell-{\tfr12}\right).\label{2.11}\end{equation}
Now if we take $y'$ from (\ref{2.9}) and $y$ from (\ref{2.11}) and
substitute them in
(\ref{2.10}), then we obtain the following equation for $m$ and $n$
\begin{equation} {(M_{\ell}'')^2\over M_{\ell}'}+4(M_{\ell}')^2
+2(xM_{\ell}'-M_{\ell})-{(2\ell-1)^2\over16M_{\ell}'}=0,\label{2.12}\end{equation}
where $m=M_{\ell}$ and $n=M_{\ell+1}$. It is easy to prove that this equation
has the Painlev\'e\ property. In
fact, equation (\ref{2.12}) is the potential version of the equation 34 of the
Gambier classification \cite{G10} (see also \cite{FA82,Ince,RG92}), which
is commonly
referred to as P34. Indeed, if we set $M_{\ell}'=Q_{\ell}$, then (\ref{2.12})
becomes
\begin{equation} Q_{\ell}''-{(Q_{\ell}')^2\over
2Q_{\ell}}+4Q_{\ell}^2+xQ_{\ell}+{(\ell-{\tfr12})^2\over
8Q_{\ell}}=0,\label{2.15}\end{equation}
which is precisely P34. Equations (\ref{2.9}) and (\ref{2.11}) can be
written as
$$\begin{array}{ll}
\displaystyle 2Q_{\ell}=y'-y^2-\tfr12{x},\qquad\qquad &\displaystyle
2Q_{\ell+1}=-y'-y^2-\tfr12{x},\\[10pt]
\displaystyle y={-2Q_{\ell}'+\ell-{\tfr12}\over
4Q_{\ell}},\qquad\qquad &\displaystyle y={-2Q_{\ell+1}'+\ell+{\tfr12}\over
4Q_{\ell+1}}.\end{array}$$
These can be interpreted as a Miura transformation between PII (\ref{1.2})
and P34 (\ref{2.15})
\cite{FA82}. According to Ramani and Grammaticos \cite{RG92}, P34
(\ref{2.15}) may be also be
thought of as a modified PII (mPII), since the relationship between their
solutions is analogous
to that between solutions of the Korteweg-de Vries and modified Korteweg-de
Vries equations (see
also \cite{FA82}).
\subsection{Auto-B\"acklund\ transformations for modified PII}
By subtracting the two equations of (\ref{2.10}), we have
\begin{equation}y(x)=M_{\ell}(x)-M_{\ell+1}(x),\label{2.18}\end{equation}
which combined with (\ref{2.11}) gives
\begin{equation} M_{\ell+1}= M_{\ell}+{4M_{\ell}''-(2\ell-1)\over 8M_{\ell}'},\qquad
M_{\ell}=M_{\ell+1}+{4M_{\ell+1}''+(2\ell+1)\over 8M_{\ell+1}'},\label{2.19}
\end{equation}
which are auto-B\"acklund\ transformations for potential mPII.
\subsection{Linear superposition for PII}
Suppose we have two solutions $M_{\ell}$ and $M_{\ell+1}$ of potential mP34
(\ref{2.12}), related by
the B\"acklund\ transformations (\ref{2.19}), then we can construct a solution of
PII (\ref{1.2}) by
using (\ref{2.18}).
\subsection{Discrete equations for potential P34}
If we let $\ell\rightarrow \ell-1$ in the second of equations (\ref{2.19}) then
\begin{equation} M_{\ell+1}= M_{\ell}+{4M_{\ell}''-(2\ell-1)\over 8M_{\ell}'},\qquad
M_{\ell-1}=
M_{\ell}+{4M_{\ell}''+(2\ell-1)\over 8M_{\ell}'}.\label{2.20}\end{equation}
Adding and subtracting these two equations gives
\begin{equation}M_{\ell}'=-\,{2\ell-1\over 4(M_{\ell+1}-M_{\ell-1})},\qquad
M_{\ell}''=-\,{(2\ell-1)(M_{\ell+1}+M_{\ell-1}-2M_{\ell})\over
4(M_{\ell+1}-M_{\ell-1})}.\label{2.21}\end{equation} By substituting
(\ref{2.21}) in (\ref{2.12}),
the result is the nonautonomous discrete equation
\begin{eqnarray*}
&&(M_{\ell-1}-M_{\ell+1})\left\{\left(\ell-{\tfr12}\right)\left[(M_{\ell-1}-M_{\ell})(M_{\
l+1}-M_{\ell})+\tfr12{x}\right]\right.\\
&&\qquad\left.
-M_{\ell}(M_{\ell-1}-M_{\ell+1})\right\}
+\left(\ell-{\tfr12}\right)^2=0,
\end{eqnarray*}
where the parameter $\ell$ can be interpreted as the discrete variable.
\setcounter{equation}{0{Painlev\'e\ IV}
In an analogous way as we studied PII in the previous section, in this
section we study PIV
(\ref{1.3}) by using the truncated expansion (\ref{1.4}) together with the
decoupling ansatz
(\ref{1.5}). The result is
\begin{eqnarray}
A&=&y+x+\tfr{1}{2}v,\label{3.2}\\
B&=&-y-x+\tfr{1}{2}r,\label{3.3}\\
0&=&(24x+6v)(A-y)-4x^2+8\ell+v^2+2(v'+y')\nonumber\\
&&\qquad-18y^2-16A^2+36yA+8AB-2rA,\label{3.4}\\
0&=&(-24x+6r)(B+y)-4x^2+8\ell+r^2+2(r'-y')\nonumber\\
&&\qquad-18y^2-16B^2-36yB+8AB-2vB.\label{3.5}
\end{eqnarray}
Equations (\ref{3.2}--\ref{3.3}) combined with (\ref{1.6}) gives $AB=k$,
with $k$ a constant. Then
substituting (\ref{3.2}--\ref{3.3}) into (\ref{3.4}--\ref{3.5}), for the
singular manifolds we
have the equations
\begin{equation} v'-\tfr12{v^2}+6k+2x^2+8\ell-2=0,\qquad
r'-\tfr12{r^2}+6k+2x^2+8\ell+2=0.\label{3.7}\end{equation}
\subsection{A Miura transformation for modified PIV}
The substitution of (\ref{2.2}-\ref{2.3}) into (\ref{1.6}) gives
\begin{equation}
y'-y^2-2xy+v'+2(k+2\ell)=0,\qquad
-y'-y^2-2xy+r'+2(k+2\ell)=0.\label{3.8}\end{equation}
It is reasonable therefore to define new functions $m$ and $n $ in the
following form
\begin{equation}
2m'=y'-y^2-2xy,\qquad 2n'=-y'-y^2-2xy.\label{3.9}\end{equation}
These equations can be integrated, taking (\ref{1.3}) into account, as
\begin{equation} \begin{array}{l}
\displaystyle 2m={{(y')^2+2\mu -(y^2+2xy)^2}\over 4y}+(2\ell+1)y,\\[10pt]
\displaystyle 2n={{(y')^2+2\mu -(y^2+2xy)^2}\over 4y}+(2\ell-1)y. \end{array}
\label{3.10}\end{equation}
Furthermore, the derivation of (\ref{3.9}) provides
\begin{equation} \begin{array}{l}
\displaystyle 2m''={2(m')^2+\mu \over y}-2(m'+2\ell+1)y,\\[10pt] \displaystyle 2n''=-{2(n')^2+\mu
\over y}+2(n'+2\ell-1)y.
\end{array}\label{3.11}\end{equation}
Taking $y'$ as defined by (\ref{3.9}) and substituting in (\ref{3.10}) we have
\begin{equation} \begin{array}{l}
\displaystyle 4(m-xm')={2(m')^2+\mu \over y}+2(m'+2\ell+1)y,\\[10pt] \displaystyle
4(n-xn')={2(n')^2+\mu \over y}+2(n'+2\ell-1)y.
\end{array}\label{3.12}\end{equation}
Adding and subtracting (\ref{3.11}) and (\ref{3.12}) the result is
\begin{equation} \begin{array}{ll}
\displaystyle y={2(m-xm')-m''\over 2(m'+2\ell+1)},\qquad &
\displaystyle{1\over y}={2(m-xm')+m''\over 2(m')^2+\mu },\\[10pt]
\displaystyle y={2(n-xn')+n''\over 2(n'+2\ell-1)},\qquad & \displaystyle
{1\over y}={2(n-xn')-n''\over 2(n')^2+\mu
}.\end{array}\label{3.13}\end{equation}
We can eliminated $y$ by multiplication of the left equation by the right
one. Thus we then have
\begin{equation}
\left(M_{\ell}''\right)^2-4\left(M_{\ell}-xM_{\ell}'\right)^2
+2\left(M_{\ell}'+2\ell+1\right)\left[2\left(M_{\ell}'\right)^2+\mu
\right]=0,\label{3.14}\end{equation}
where
$m=M_{\ell}$ and $n=M_{\ell-1}$. Equation (\ref{3.14}) is of
Painlev\'e\ type, and it can be considered as the modified version of PIV (mPIV). The
Miura transformation that relates PII and mPII are (\ref{3.11}), which can
now be written as
\begin{equation}
2M_{\ell}'=y'-y^2-2xy,\qquad 2M_{\ell-1}'=-y'-y^2-2xy.\label{3.15}\end{equation}
\subsection{Auto-B\"acklund\ transformations}
Subtracting the two equations of (\ref{3.10}), and with the aid of
(\ref{3.15}), we have
\begin{equation} y=M_{\ell}-M_{\ell-1},\label{3.17}\end{equation}
then combined with the left part of (\ref{3.13}) yields to
\begin{equation} \begin{array}{l}
\displaystyle M_{\ell-1}=M_{\ell}+{M_{\ell}''-2(M_{\ell}-xM_{\ell}')\over
2(M_{\ell}'+2\ell+1)},\\[10pt]
\displaystyle M_{\ell}=M_{\ell-1}+{M_{\ell-1}''+2(M_{\ell-1}-xM_{\ell-1}')\over
2(M_{\ell-1}'+2\ell-1)},
\end{array}\label{3.18}\end{equation}
which are auto-B\"acklund\ transformation for mPIV (\ref{3.14}).
\subsection{Linear superposition for PIV}
If we have two solutions $M_{\ell}$ and $M_{\ell+1}$ of potential mPIV
(\ref{3.14}),
related by the B\"acklund\ transformations (\ref{3.18}), then we can construct a
solution of PIV
(\ref{1.3}) by using (\ref{3.17}).
\subsection{Discrete equations for mPIV}
As we did in the previous section, the B\"acklund\ transformations (\ref{3.18})
can be written as
\begin{equation} \begin{array}{l}
\displaystyle M_{\ell-1}=M_{\ell}+{M_{\ell}''-2(M_{\ell}-xM_{\ell}')\over
2(M_{\ell}'+2\ell+1)},\\[10pt]
\displaystyle M_{\ell+1}=M_{\ell} +{M_{\ell}''+2(M_{\ell}-xM_{\ell}')\over 2(M_{\ell}'+2\ell+1)}.
\end{array}\label{3.19}\end{equation}
Then by addition and subtraction, it is easy to show that
\begin{equation} \begin{array}{l} \displaystyle
M_{\ell}'={2M_{\ell}-(2\ell+1)(M_{\ell+1}-M_{\ell-1})\over
M_{\ell+1}-M_{\ell-1}+2x},\\[10pt] \displaystyle
M_{\ell}''={(M_{\ell}'+2\ell+1)(M_{\ell+1}+M_{\ell-1}-2M_{\ell})\over
M_{\ell+1}-M_{\ell-1}+2x},\end{array}\label{3.20}\end{equation}
whose substitution in (\ref{3.15}) yields the discrete equation
\begin{eqnarray*}
&&(M_{\ell+1}-M_{\ell-1}+2x)(M_{\ell}-M_{\ell-1})(M_{\ell}-M_{\ell+1})\left[M_{\ell}+2\left(\ell
+{\tfr12}\right)x\right]\\
&&\qquad+\left\{2\left[M_{\ell}-\left(\ell+{\tfr12}\right)\right]
(M_{\ell+1}-M_{\ell-1})\right\}^2
+\tfr12{x}\left(M_{\ell+1}-M_{\ell-1}+2x\right)^2=0.
\end{eqnarray*}
where $\ell$ is the discrete parameter.
\setcounter{equation}{0{Conclusions}
In this paper we have derived Miura transformations, modified equations and
associated discrete
equations for the second and fourth Painlev\'e\ equations. Recently there have been
several studies of
the derivation of discrete equations, and in particular the discrete Painlev\'e\
equations, from B\"acklund\
transformations of the (continuous) Painlev\'e\ equations
\cite{refFGR,refGNR,refGra,GR98,refGTii,refGTiii,NSKGR96,refTsegb,refTsegc}.
Hierarchies of solutions of PII and PIV are well-known (cf.\
\cite{refAirault,BCH95,PAC90,refGromaki,Gromak,refLuka,Murata,refOka,refUW}).
Since there is an explicit relationship between PII and PIV and discrete
equations, then these Hierarchies of
solutions of PII and PIV also satisfy difference equations in addition to
the ordinary
differential equations. This is analogous to the situation for the
classical special functions,
such as Bessel, hypergeometric, Legendre, Weber-Hermite and Whittaker
functions, which satisfy
both an ordinary differential equation and a recurrence relation, which is
a discrete equation.
This provides further evidence that the Painlev\'e\ equations may be thought of as
nonlinear special
functions and that there is a deep relationship between the classical
special functions, the Painlev\'e\
equations and the discrete Painlev\'e\ equations (see, for example,
\cite{refTRGK}).
\def\frenchspacing\it{\frenchspacing\it}
|
\section*{Introduction}
Our introduction is rather brief since this paper mainly is a
continuation of \cite{prep:wir}. There we a gave
a formula for a Lorentz invariant renormalization in one
coordinate. We use the same methods to construct a covariant solution
for more variables.
We review the EG subtraction and give some useful formulas in
section~\ref{sec:sub}. The cohomological analysis that leads to a group
covariant solution is reviewed in section~\ref{sec:g}. In
sections~\ref{sec:tensor},~\ref{sec:spinor} we give an inductive
construction for these solutions in the case of tensorial and
spinorial Lorentz covariance. Since the symmmetry of the one variable
problem is absent in general some permutation group calculus will
enter. Necessary material is in the appendix. In the last
section~\ref{sec:bphz} we give a covariant BPHZ subtraction for
arbitrary (totally spacelike) momentum by fourier transformation.
\section{The EG subtraction}
\label{sec:sub}
We review the extension procedure in the EG approach \cite{pap:ep-gl}.
A more general introduction can be found in \cite{pap:prange1}, a
generalization to manifolds in \cite{proc:fred-brun,prep:fred-brun2}.
Let $ \ensuremath{\mathcal{D}}^{\omega}(\Rd) $ be the subspace of test functions
vanishing to order $ \omega $ at $ 0 $. As usual, any operation on
distributions is defined by the corresponding action on testfunctions.
Define
\begin{gather}
\W{\omega}{w}:\ensuremath{\mathcal{D}}(\Rd)\rightarrow\ensuremath{\mathcal{D}}^{\omega}(\Rd),
\quad\varphi\rightarrow\W{\omega}{w}\varphi, \notag \\
\left(\W{\omega}{w}\varphi\right)(x)=\varphi(x)-w(x)\sum_{|\alpha|\leq\omega}
\frac{x^{\alpha}}{\alpha!}\ensuremath{\partial}_{\alpha}\left(\varphi w^{-1}\right)(0),
\label{def:W}
\end{gather}
with $ w\in\ensuremath{\mathcal{D}}(\Rd), w(0)\not=0 $. If ${^0t}$ is a distribution on
$\ensuremath{\mathcal{D}}(\Rdon)$ with singular order $ \omega $ then ${^0t}$ can be
defined uniquely on $ \ensuremath{\mathcal{D}}^{\omega}(\Rd) $, and
\begin{equation}
\scp{\tR{\omega}{w}}{\varphi}\doteq\scp{{^0t}}{\W{\omega}{w}\varphi}
\label{def:tR}
\end{equation}
defines an extension -- called renormalization -- of $^0t $ to the
whole testfunction space, $ \tR{\omega}{w}\in\ensuremath{\mathcal{D}}'(\Rd) $. With the Leibnitz rule
\begin{equation}
\ensuremath{\partial}_{\gamma}(fg)
=\gamma!\sum_{\mu+\nu=\gamma}\frac{1}{\mu!\nu!}\ensuremath{\partial}_{\mu}f\,\ensuremath{\partial}_{\nu}g,
\label{eq:leibnitz}
\end{equation}
we find
\begin{equation}
\ensuremath{\partial}_{\gamma}\biggl(w\sum_{|\alpha|\leq\omega}
\frac{x^{\alpha}}{\alpha!}\ensuremath{\partial}_{\alpha}(\varphi w^{-1})(0)\biggr)(0)
=\ensuremath{\partial}_{\gamma}\varphi(0),
\label{eq:diffrest}
\end{equation}
for $ |\gamma|\leq\omega $, verifying the projector properties of
$W$. The $W$-operation (\ref{def:W}) is simplified if we require
$w(0)=1$ and $\partial_{\alpha}w(0)=0$, for $0<|\alpha|\leq\omega$
(this was our assumption in \cite{prep:wir}). It can be achieved by
the following $V$-operation ($\ensuremath{\partial}_\mu w^{-1}$ means
$\ensuremath{\partial}_{\mu}(w^{-1})$):
\begin{equation}
V_{\omega}:\ensuremath{\mathcal{D}}(\Rd)\mapsto\ensuremath{\mathcal{D}}(\Rd),\quad (V_{\omega}w)(x)\doteq
w(x)\sum_{|\mu|\leq \omega}\frac{x^{\mu}}{\mu!}\ensuremath{\partial}_{\mu}w^{-1}(0),
\label{def:V}
\end{equation}
where $w(0)\not=0$ is still assumed. We can write $W$ as
\begin{equation}
\label{eq:varW}
\left(\W{\omega}{w}\varphi\right)(x)
=\varphi(x)-\sum_{|\alpha|\leq\omega}
\frac{x^{\alpha}}{\alpha!}V_{\omega-|\alpha|}w\,\ensuremath{\partial}_{\alpha}\varphi(0).
\end{equation}
The extension \eqref{def:tR} is not unique. We can add any polynomial
in derivatives of $\delta$ up to order $\omega$:
\begin{align}
\scp{\tR{\omega}{w}}{\varphi}&=\scp{\tiltR{\omega}{w}}{\varphi}
+\sum_{|\alpha|\leq\omega}
\frac{a^{\alpha}}{\alpha!}\ensuremath{\partial}_{\alpha}\varphi(0), \\
\intertext{or rearranging the coefficients}
&=\scp{\tiltR{\omega}{w}}{\varphi}
+\sum_{|\alpha|\leq\omega}
\frac{c^{\alpha}}{\alpha!}\ensuremath{\partial}_{\alpha}\left(\varphi w^{-1}\right)(0)
\label{def:tiltR}
\end{align}
Since $\W{\omega}{w}( w x^\alpha) = \W{\omega}{w} (x^\alpha
V_{\omega-|\alpha|}w) = 0$ for $|\alpha| \leq \omega$, $c$ resp.\ $a$
are given by
\begin{align}
a^\alpha&=\scp{\tiltR{\omega}{w}}{x^\alpha V_{\omega-|\alpha|}w}, &
c^\alpha&=\scp{\tiltR{\omega}{w}}{x^\alpha w}.
\label{eq:cundaaust}
\end{align}
They are related through:
\begin{align}
a^\alpha&=c^\alpha\sum_{|\mu|\leq\omega-|\alpha|}
\frac{c^{\mu}}{\mu!}\ensuremath{\partial}_{\mu}w^{-1}(0), &
c^\alpha&=a^\alpha\sum_{|\mu|\leq\omega-|\alpha|}
\frac{a^{\mu}}{\mu!}\ensuremath{\partial}_{\mu}w(0), &
1\leq|\alpha|\leq\omega, \notag \\
a^0&=\sum_{|\mu|\leq\omega}
\frac{c^{\mu}}{\mu!}\ensuremath{\partial}_{\mu}w^{-1}(0), &
c^0&=\sum_{|\mu|\leq\omega} \frac{a^{\mu}}{\mu!}\ensuremath{\partial}_{\mu}w(0).
\label{eq:causa}
\end{align}
The equation for $a$ follows from the Leibnitz rule in
\eqref{def:tiltR}, while the equation for $c$ is derived from
\eqref{eq:cundaaust}.
In quantum field theory the coefficients $a$ are called counter terms.
They are not arbitrary. They have to be chosen in such a way that
Lorentz covariance of $^{0}t$ is preserved in the extension. This
follows in the next sections. The remaining freedom is further
restricted by discrete symmetries like permutation symmetry or $C,P$
and $T$ symmetries. At the end gauge invariance or renormalization
constraints will fix the extension uniquely. But up to now there is no
local prescription of the latter.
\section{The $G$-covariant extension}
\label{sec:g}
We will first define the notion of a $G$-covariant distribution. So
let $G$ be a linear transformation group on \Rd\ i.e. $x\mapsto gx$,
$g\in G$. Then
\begin{equation}
x^{\alpha}\mapsto{g^{\alpha}}_{\beta}x^{\beta}=(gx)^{\alpha}
\label{}
\end{equation}
denotes the corresponding tensor representation. $G$ acts on
functions in the following way:
\begin{equation}
(g\varphi)(x)\doteq\varphi(g^{-1}x),
\label{def:gphi}
\end{equation}
so that \ensuremath{\mathcal{D}}\ is made a $G$-module. We further have
\begin{align}
g(\varphi\psi)&=(g\varphi)(g\psi),\label{eq:gprod} \\
x^{\alpha}\ensuremath{\partial}_{\alpha}(g^{-1}\varphi)
&=(gx)^{\alpha}g^{-1}(\ensuremath{\partial}_{\alpha}\varphi), \label{eq:1} \\
x^{\alpha}\ensuremath{\partial}_{\alpha}(g^{-1}\varphi)(0)
&=(gx)^{\alpha}\ensuremath{\partial}_{\alpha}\varphi(0). \label{eq:2}
\end{align}
Now assume we have a distribution ${^0t}\in\ensuremath{\mathcal{D}}'(\Rdon)$ that transforms
covariantly under the Group $G$ as a density, i.e.
\begin{equation}
{^0t}(gx)|\det g|=D(g){^0t}(x),
\label{eq:tcov}
\end{equation}
where $D$ is the corresponding representation. That means:
\begin{equation}
\scp{{^0t}}{g\psi}=\scp{D(g){^0t}}{\psi}\doteq D(g)\scp{{^0t}}{\psi}.
\label{def:tcov}
\end{equation}
We will now investigate what happens to the covariance in the
extension process. We compute:
\begin{align}
D(g)&\scp{\tR{\omega}{w}}{g^{-1}\varphi}
-\scp{\tR{\omega}{w}}{\varphi}
\label{eq:cov1} \\
&=D(g)\scp{{^0t}}{\W{\omega}{w}g^{-1}\varphi}
-\scp{{^0t}}{\W{\omega}{w}\varphi}
\label{eq:cov2} \\
&\stackrel{\eqref{eq:varW}}{=}
D(g)\scp{{^0t}}{g^{-1}\varphi-\sum_{|\alpha|\leq\omega}
\frac{x^{\alpha}}{\alpha!}V_{\omega-|\alpha|}w\,\ensuremath{\partial}_{\alpha}(g^{-1}\varphi)(0)}
-\scp{{^0t}}{\W{\omega}{w}\varphi} \label{eq:cov3} \\
&\stackrel{(\ref{eq:gprod},\ref{eq:2})}{=}
D(g)\scp{{^0t}}{g^{-1}\biggl(\varphi- \sum_{|\alpha|\leq\omega}
\frac{x^{\alpha}}{\alpha!}\left(gV_{\omega-|\alpha|}w\right)
\ensuremath{\partial}_{\alpha}\varphi(0)\biggr)}
-\scp{{^0t}}{\W{\omega}{w}\varphi}
\label{eq:cov4} \displaybreak[0]\\
&\stackrel{(\ref{def:tcov})}{=}
\sum_{|\alpha|\leq\omega}
\scp{{^0t}}{x^{\alpha}(\eins-g)(V_{\omega-|\alpha|}w)}
\frac{\ensuremath{\partial}_{\alpha}\varphi(0)}{\alpha!} \label{eq:cov8} \\
&\doteq\sum_{|\alpha|\leq\omega}b^{\alpha}(g)
\frac{\ensuremath{\partial}_{\alpha}\varphi(0)}{\alpha!}.
\label{def:balpha}
\end{align}
Then (\ref{def:balpha}) defines a map from $G$ to a finite dimensional
complex vectorspace. Now we follow \cite{prep:stora-pop},
\cite{bk:scharf}[chapter 4.5]: Applying two transformations
\begin{align}
b^{\alpha}(g_{1}g_{2})
&=\scp{{^0t}}{x^{\alpha}(\eins-g_{1}g_{2})(V_{\omega-|\alpha|}w)} \\
&=\scp{{^0t}}{x^{\alpha}\left((\eins-g_{1})
+g_{1}(\eins-g_{2})\right)(V_{\omega-|\alpha|}w)} \\
&=b^{\alpha}(g_{1})+|\det g_{1}|\scp{{^0t}(g_{1}x)}
{(g_{1}x)^{\alpha}(\eins-g_{2})(V_{\omega-|\alpha|}w)},
\end{align}
and omitting the indices we see
$b(g_{1}g_{2})=b(g_{1})+D(g_{1})g_{1}b(g_{2}), $ which is a 1-cocycle
for $ b(g) $. Its trivial solutions are the 1-coboundaries
\begin{equation}
b(g)=(\eins-D(g)g)a,
\label{def:cobound}
\end{equation}
and these are the only ones if the first cohomology group of $G$ is
zero. In that case we can restore $G$-covariance by adding the
following counter terms:
\begin{equation}
\scp{\tR{\omega}{w}^{G-\mathrm{cov}}}{\varphi}
\doteq\scp{\tR{\omega}{w}}{\varphi}
+\sum_{|\alpha|\leq\omega}
\frac{1}{\alpha!}a^{\alpha}(w)\ensuremath{\partial}_{\alpha}\varphi(0).
\label{def:tRGcov}
\end{equation}
The task is to determine $a$ from (\ref{def:cobound}) and (\ref{eq:cov8}):
\begin{equation}
\scp{{^0t}}{x^{\alpha}(\eins-g)(V_{\omega-|\alpha|}w)}
=\bigl[(\eins-D(g)g)a\bigr]^\alpha
\label{def:a}
\end{equation}
\section{Tensorial Lorentz covariance}
\label{sec:tensor}
The first cohomology group of $\Lcal_{+}^{\uparrow}$ vanishes
\cite{bk:scharf}[chapter 4.5 and references there]. We determine $ a $
from the last equation. The most simple solution appears in the case
of Lorentz invarinance in one coordinate. This situation was
completely analyzed in \cite{prep:wir} for $\ensuremath{\partial}_{\alpha}w(0) =
\delta_{\alpha}^{0}$. The following two subsections generalize the
results to arbitrary $w,\ w(0)\not=0$.
\subsection{Lorentz invariance in \Rv}
\label{subsec:Linv4d}
If we expand the index $\alpha$ into Lorentz indices $\mu_1, \dots,
\mu_n$, \eqref{def:a} is symmetric in $\mu_1, \dots, \mu_n$ and
therefore $a$ will be, too. We just state our result from
\cite{prep:wir} which is modified through the generalization for the
choice of $w$:
\begin{multline}
a^{(\mu _1 \ldots \mu_n)} = \frac{ (n-1)!!}{(n+2)!!}
\sum_{s=0}^{\left[\frac{n-1}{2} \right] }
\frac{(n-2s)!!}{(n-2s-1)!!} \eta^{(\mu_1 \mu _2}
\ldots \eta^{\mu_{2s-1}\mu _{2s}} \times \\
\times \left\langle
{^0t}, (x^2)^s x ^{\mu _{2s+1}} \ldots x ^{\mu_{n-1}}
\left(x^2\ensuremath{\partial}^{\mu_{n})}-x^{\mu_{n})}x^{\beta}\ensuremath{\partial}_{\beta}\right)
V_{\omega-n}w\right\rangle,
\label{eq:amun}
\end{multline}
if we choose the fully contracted part of $ a $ to be zero in case
of $ n $ being even. We used the notation
\begin{align*}
b^{(\mu_{1}\dots\mu_{n})}&=\frac{1}{n!}\sum_{\pi\in S_{n}}
b^{\mu_{\pi(1)}\dots\mu_{\pi(n)}}, &
b^{[\mu_{1}\dots\mu_{n}]}&=\frac{1}{n!}\sum_{\pi\in S_{n}}
\sgn(\pi)b^{\mu_{\pi(1)}\dots\mu_{\pi(n)}},
\end{align*}
for the totally symmetric resp.\ antisymmetric part of a tensor.
\subsection{Dependence on $w$}
\label{subsubsec:scale}
Performing a functional derivation of the Lorentz invariant extension
with respect to $w$, only Lorentz invariant counter terms appear.
\begin{definition}
The functional derivation is given by:
\begin{equation*}
\scp{\frac{\delta}{\delta g}F(g)}{\psi}\doteq
\left.\frac{\dif}{\dif\lambda}F(g+\lambda\psi)\right\vert_{\lambda=0}.
\end{equation*}
\end{definition}
This definition implies the following functional derivatives:
\begin{align}
\scp{\frac{\delta}{\delta w}\tR{\omega}{w}(\varphi)}{\psi}
&=-\sum_{|\alpha|\leq\omega}\frac{1}{\alpha!}
\scp{\tR{\omega}{w}}{x^\alpha\psi}
\ensuremath{\partial}_\alpha\left(\varphi w^{-1}\right)(0),\label{eq:derivtR} \\
\scp{\frac{\delta}{\delta w}\scp{S}{V_\omega w}}{\psi}
&=\sum_{|\alpha|\leq\omega}\frac{1}{\alpha!}
\scp{S}{\W{\omega}{w}(x^\alpha\psi)}
\ensuremath{\partial}_\alpha w^{-1} (0), \label{eq:derivS}
\end{align}
for any distribution $S$.
\begin{proof}
We show how to derive the first relation. Inserting the definition we
find:
\begin{align*}
\left.\frac{\dif}{\dif\lambda}
\tR{\omega}{w+\lambda\psi}(\varphi)\right\vert_{\lambda=0}
&=\scp{^0t}{-\psi\sum_{|\alpha|\leq\omega}
\frac{x^\alpha}{\alpha!}\ensuremath{\partial}_\alpha\left(\varphi w^{-1}\right)(0)
+w\sum_{|\alpha|\leq\omega}\frac{x^\alpha}{\alpha!}
\ensuremath{\partial}_\alpha\left(\varphi\psi w^{-2}\right)(0)}, \\
\intertext{using Leibnitz rule and rearranging the summation in the
second term,}
&\begin{aligned}
=\sum_{|\alpha|\leq\omega}\frac{1}{\alpha!}
\Biggl\langle{^0t},-&\psi x^\alpha\ensuremath{\partial}_\alpha\left(\varphi w^{-1}\right)(0) \\
+&w x^\alpha\ensuremath{\partial}_\alpha\left(\varphi w^{-1}\right)(0)
\sum_{|\nu|\leq\omega-|\alpha|}\frac{x^\nu}{\nu!}
\ensuremath{\partial}_\nu\left(\psi w^{-1}\right)(0) \Biggr\rangle
\end{aligned} \\
&=-\sum_{|\alpha|\leq\omega}\frac{1}{\alpha!}
\scp{^0t}{x^\alpha\W{\omega-|\alpha|}{w}\psi}
\ensuremath{\partial}_\alpha\left(\varphi w^{-1}\right)(0) \\
&=-\sum_{|\alpha|\leq\omega}\frac{1}{\alpha!}
\scp{\tR{\omega}{w}}{x^\alpha\psi}\ensuremath{\partial}_\alpha\left(\varphi w^{-1}\right)(0),
\end{align*}
where we used the relation $x^\alpha\W{\omega-|\alpha|}{w}\varphi =
\W{\omega}{w}(x^\alpha\varphi)$ on the last line. The second equation
follows from a similar calculation.
\end{proof}
We calculate the dependce of $a$ on $w$. With \eqref{eq:derivS} we
get:
\begin{multline*}
\frac{\delta}{\delta w} a^{(\mu_1\dots\mu_n)}(w)
= \frac{(n-1)!!}{(n+2)!!}
\sum_{s=0}^{\left[\frac{n-1}{2} \right] }
\frac{(n-2s)!!}{(n-1-2s)!!} \eta^{(\mu_1\mu_2} \dots
\eta^{\mu _{2s-1} \mu _{2s}} \times \\
\times\sum_{|\beta|\leq\omega-n}\frac{1}{\beta!}
\scp{^0t}{(x^2)^s x^{\mu_{2s+1}} \dots x ^{\mu_{n-1}}
\left(x^2\ensuremath{\partial}^{\mu _{n})}
-x^{\mu_{n})}x^{\beta}\ensuremath{\partial}_{\beta}\right)\W{\omega-n}{w}(x^\beta\psi)}
\ensuremath{\partial}_\beta w^{-1}(0).
\label{eq:damun}
\end{multline*}
To condense the notation we again use $\beta$ as a multiindex. Since
$\W{\omega-n}{w}(x^\beta\psi)$ is sufficient regular, we can put the
$x$'s and derivatives on the left and the same calculation like in
\cite{prep:wir} applies. The result is
\begin{multline*}
\scp{\frac{\delta}{\delta w}a^{\mu_{1}\cdots\mu_{n}}(w)}{\psi}=
\sum_{|\beta|\leq\omega-n}\frac{\ensuremath{\partial}_\beta w^{-1}(0)}{\beta!}
\Biggl[
\scp{\tR{\omega}{w}}{x^{\mu_{1}}\cdots x^{\mu_{n}}x^\beta\psi}
+ \\
-\begin{cases}
0,&n\text{ odd}, \\
\frac{2(n-1)!!}{(n+2)!!}
\scp{\tR{\omega}{w}}{(x^2)^\frac{n}{2}x^\beta\psi}
\eta^{(\mu_1\mu_2}\cdots\eta^{\mu_{n-1}\mu_n)},
&n\text{ even.}
\end{cases}
\Biggr]
\end{multline*}
Using this result and \eqref{eq:derivtR} we find:
\begin{align*}
\scp{\frac{\delta}{\delta w}\scp{\tRli{\omega}{w}}{\phi}}{\psi}
&=-\sum_{\substack{n=0\\ n\text{ even}}}^{\omega}\frac{d_{n}}{n!}
\square^{\frac{n}{2}}\phi(0), \\
d_{n}&\doteq
\frac{2(n-1)!!}{(n+2)!!}\sum_{|\beta|\leq\omega-n}\frac{1}{\beta!}
\scp{\tR{\omega}{w}}{(x^{2})^{\frac{n}{2}}x^\beta\psi}
\ensuremath{\partial}_\beta w^{-1}(0),
\end{align*}
where we set $ d_{0}=1 $.
\subsection{General Lorentz covariance}
If the distribution ${^{0}t}$ depends on more than one variable,
${^{0}t}x^{\alpha}$ will not be symmetric in all Lorentz indices in general.
Since $x^{\alpha}$ transforms like a tensor, it is natural to
generalize the discussion to the case, where ${^{0}t}$ transforms like a
tensor, too. Assume rank$( {^0t} )=r$, then $D(g)g$ is the tensor
representation of rank $p=r+n,n=|\alpha|$, in (\ref{def:a}). From now
on we will omit the indices. So if $t\in\ensuremath{\mathcal{D}}(\Rvmon)$, we denote by
\xbar\ -- formerly $x^{\alpha}$ -- a tensor of rank $n$ built of
$x_{1},\dots,x_{m}$.
To solve \eqref{def:a} we proceed like in
\cite{prep:wir}. Since the equation holds for all $g$ we will solve
for $a$ by using Lorentz transformations in the infinitesimal
neighbourhood of $\eins$. If we take
$\theta_{\alpha\beta}=\theta_{[\alpha\beta]} $ as six coordinates these transformations read:
\begin{equation}
g \approx\eins+\frac{1}{2}\theta_{\alpha\beta}l^{\alpha\beta},
\label{eq:gbyone}
\end{equation}
with the generators
\begin{equation}
{(l^{\alpha\beta})^{\mu}}_{\nu}
=\eta^{\alpha\mu}\delta^{\beta}_{\nu}-\eta^{\beta\mu}\delta^{\alpha}_{\nu}.
\label{def:generator}
\end{equation}
Then, for an infinitesimal transformation one
finds from \eqref{def:a}:
\begin{gather}
B^{\alpha\beta}\doteq
2\biggl\langle{^0t},\xbar\sum_{j=1}^{m}
x^{[\alpha}_j\ensuremath{\partial}^{\beta]}_j(V_{\omega-n}w)\biggr\rangle
=(l^{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\eins\otimes\dots\otimes l^{\alpha\beta})a,
\label{eq:atensor}
\end{gather}
$\alpha,\beta$ being Lorentz four-indices. In \cite{prep:wir}
our ability to solve that equation heavily relied on the given
symmetry, which is in general absent here. Nevertheless we can find
an inductive construction for $a$, corresponding to equation
(29) in \cite{prep:wir}.
We build one Casimir operator on the r.h.s. (the other one is always
zero, since we are in a $(1/2,1/2)^{\otimes p}$ representation).
\subsubsection*{The case $p=1$}
Just to remind that $p$ is the rank of $\xbar t$, this occurs if
either $t$ is a vector and $\xbar=1, (n=0)$, or $t$ is a scalar and
$\xbar=x_{1},\dots,x_{m}$. (\ref{eq:atensor}) gives:
\begin{equation}
\frac{1}{2}l_{\alpha\beta}B^{\alpha\beta}
=\frac{1}{2}l_{\alpha\beta}l^{\alpha\beta}a
=-3\eins a,
\label{eq:cas1}
\end{equation}
since the Casimir operator is diagonal in the irreducible $(1/2,1/2)$
representation.
\subsubsection*{The case $p=2$}
We get
\begin{equation}
\frac{1}{2}(l_{\alpha\beta}\otimes\eins
+\eins\otimes l_{\alpha\beta})B^{\alpha\beta}
=(-6\eins+l_{\alpha\beta}\otimes l^{\alpha\beta})a.
\label{eq:acas2}
\end{equation}
Since $a$ is a tensor of rank 2, let us introduce the projector onto the
symmetric resp. antisymmetric part and the trace:
\begin{align}
{P_{S}^{\mu\nu}}_{\rho\sigma}
&=\frac{1}{2}(\delta^{\mu}_{\rho}\delta^{\nu}_{\sigma}
+\delta^{\nu}_{\rho}\delta^{\mu}_{\sigma}), &
{P_{A}^{\mu\nu}}_{\rho\sigma}
&=\frac{1}{2}(\delta^{\mu}_{\rho}\delta^{\nu}_{\sigma}
-\delta^{\nu}_{\rho}\delta^{\mu}_{\sigma}), \label{def:P} &
{P_\eta^{\mu\nu}}_{\rho\sigma}
&=\frac{1}{4}\eta^{\mu\nu}\eta_{\rho\sigma}, \\
P^2&=P, &
P_S+P_A&=\eins, &
P_S-P_A&=\tau,
\end{align}
where $\tau$ denotes the permutation of the two indices.
Using (\ref{def:generator}), we find
\begin{equation}
\frac{1}{2}l_{\alpha\beta}\otimes l^{\alpha\beta}=4P_{\eta}-\tau.
\label{eq:genten}
\end{equation}
Now we insert (\ref{eq:genten}) into (\ref{eq:acas2}). The trace part
will be set to zero again. Acting with $P_A$ and $P_S$ on the
resulting equation gives us two equations for the antisymmetric and
symmetric part respectively. This yields:
\begin{equation}
a=-\frac{1}{16}(P_S+2P_A)(l_{\alpha\beta}\otimes\eins
+\eins\otimes l_{\alpha\beta})B^{\alpha\beta}.
\label{eq:a2}
\end{equation}
\subsubsection*{Inductive assumption}
Now we turn back to equation \eqref{def:a}. We note that any
contraction commutes with the (group) action on the rhs. Hence, if we
contract \eqref{eq:atensor}, we find on the rhs:
\begin{multline*}
\eta_{ij}(l^{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\eins\otimes\dots\otimes l^{\alpha\beta})a= \\
(l^{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\widehat{i}+\dots+\widehat{j}+\dots
+\eins\otimes\dots\otimes l^{\alpha\beta})(\eta_{ij}a),
\end{multline*}
where $i,j$ denote the positions of the corresponding indices, and the
$\widehat{\text{hat}}$ means omission. Therefore the rank of
\eqref{eq:atensor} is reduced by two and we can proceed inductively.
With the cases $p=1, p=2$ solved, we assume that all possible
contractions of $a$ are known.
\subsubsection*{Induction step}
Multiplying (\ref{eq:atensor}) with the generator and contracting the
indices yields:
\begin{multline}
\biggl(3p\eins+2\sum_{\tau\in S_p}\tau\biggr)a
=-\frac{1}{2}(l_{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\eins\otimes\dots\otimes l_{\alpha\beta})B^{\alpha\beta}
+8\sum_{i<j\leq p}{P_\eta}_{ij}a.
\label{eq:acasp}
\end{multline}
The transposition $\tau$ acts on $a$ by permutation of the
corresponding indices. For a general $\pi\in S_p$ the action on $a$ is
given by: $\pi
a^{\mu_1\dots\mu_p}=a^{\mu_{\pi^{-1}(1)}\dots\mu_{\pi^{-1}(p)}}$. In
order to solve this equation we consider the representation of the
symmetric groups. We give a brief summary of all necessary ingredients
in appendix~\ref{app:permrep}. So let $k_\tau\doteq\sum_{\tau\in
S_p}\tau$ be the sum of all transpositions of $S_p$. Then $k_\tau$ is
in the center of the group algebra $\Acal_{S_p}$. It can be
decomposed into the idempotents $e_{(m)}$ that generate the irreducible
representations of $S_p$ in $\Acal_{S_p}$.
\begin{equation}
k_\tau=h_\tau\sum_{(m)}\frac{1}{f_{(m)}}\chi_{(m)}(\tau)e_{(m)}.
\label{eq:decompktau}
\end{equation}
The sum runs over all partitions $(m)=(m_1,\dots,m_r), \sum_{i=1}^r
m_i=p, m_1\geq m_2\geq\dots\geq m_r$ and $h_\tau=\frac{1}{2} p(p-1)$
is the number of transpositions in $S_p$. $\chi_{(m)}$ is the
character of $\tau$ in the representation generated by $e_{(m)}$ which
is of dimension $f_{(m)}$. We use \eqref{eq:decompktau}, the
ortogonality relation $e_{(m)}e_{(m')} =\delta_{(m)(m')}$ and the
completeness $\sum_{(m)}e_{(m)}=\eins$ in (\ref{eq:acasp}). The
expression in brackets on the l.h.s may be orthogonal to some
$e_{(m)}$. The corresponding $e_{(m)}a$ contribution will be any
combinations of $\eta$'s and $\epsilon$'s --$\epsilon$ being the
totally antisymmetric tensor in four dimensions -- transforming
correctly and thus can be set to zero. We arrive at
\begin{multline}
a=\sum_{\substack{(m)\\c(m)\not=0}}\frac{e_{(m)}}{c(m)}
\left(
-\frac{1}{2}(l_{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\eins\otimes\dots\otimes l_{\alpha\beta})B^{\alpha\beta}
+8\sum_{i<j\leq p}{P_\eta}_{ij}a
\right), \\
c(m)\doteq 3p+p(p-1)\frac{\chi_{(m)}(\tau)}{f_{(m)}}
=3p+\sum_{i=1}^r\left(
b_i^{(m)}(b_i^{(m)}+1)-a_i^{(m)}(a_i^{(m)}+1)
\right),
\label{eq:atenrec}
\end{multline}
with $a=(a_1,\dots,a_r), b=(b_1,\dots,b_r)$ denoting the
characteristics of the frame $(m)$, see appendix~\ref{app:permrep}.
Let us take $p=4$ as an example:
\[
\begin{array}{cccc}
\text{idempotent} & \text{Young frame} & \text{dimension} &
\text{character} \\
& & & \\
e_{(4)} & \yng(4) & f_{(4)}=1 & \chi_{(4)}(\tau)=1 \\
& & & \\
e_{(3,1)} & \yng(3,1) & f_{(3,1)}=3 & \chi_{(3,1)}(\tau)=1 \\
& & & \\
e_{(2,2)} & \yng(2,2) & f_{(2,2)}=2 & \chi_{(2,2)}(\tau)=0 \\
& & & \\
e_{(2,1,1)} & \yng(2,1,1) & f_{(2,1,1)}=3 & \chi_{(2,1,1)}(\tau)=-1 \\
& & & \\
e_{(1,1,1,1)} & \yng(1,1,1,1) & f_{(1,1,1,1)}=1 & \chi_{(1,1,1,1)}(\tau)=-1
\end{array}
\]
We find for (\ref{eq:acasp})
\begin{equation}
a=\frac{1}{48}(2e_{(4)} + 3e_{(3,1)} +4e_{(2,2)} +6e_{(2,1,1)})\times
\text{r.h.s}(\ref{eq:acasp}).
\end{equation}
We see that no $e_{(1,1,1,1)}$ appears in that equation. It corresponds to the one dimensional
$sgn$-representation of $S_4$, so $e_4a\propto\epsilon$.
\section{Spinorial Lorentz covariance}
\label{sec:spinor}
In this section we follow \cite{bk:sexl-urb}. The most general
representation of $\Lcal_{+}^{\uparrow}$ can be built of tensor
products of \ensuremath{\mathit{SL}(2,\menge{C})}\ and $\overline{\ensuremath{\mathit{SL}(2,\menge{C})}}$ and direct sums of these. A
two component spinor $\Psi$ transforms according to
\begin{equation}
\Psi^{A}={g^{A}}_{B}\Psi^{B},
\label{eq:spintrafo}
\end{equation}
where $g$ is a $2\times2$-matrix in the \ensuremath{\mathit{SL}(2,\menge{C})}\ representation of
$\Lcal_{+}^{\uparrow}$. For the complex conjugated representation we
use the dotted indices, i.e.
\begin{equation}
\overline{\Psi}^{\Xdot}={\overline{g}^{\Xdot}}_{\Ydot}\overline{\Psi}^{\Ydot},
\label{eq:ccspintrafo}
\end{equation}
with $ {\overline{g}^{\Xdot}}_{\Ydot}=\overline{{g^{X}}_{Y}} $ in the $
\overline{\ensuremath{\mathit{SL}(2,\menge{C})}} $ representation. The indices are lowered and raised
with the $ \epsilon $-tensor.
\begin{gather}
\epsilon_{AB}=\overline{\epsilon}_{\dot{A}} \newcommand{\Bdot}{\dot{B}\Bdot}
\doteq\epsilon_{\dot{A}} \newcommand{\Bdot}{\dot{B}\Bdot}, \label{eq:eps} \\
\epsilon^{AB}\epsilon_{AC}
=\epsilon^{BA}\epsilon_{CA}
=\delta^{B}_{C}.
\end{gather}
We define the Van-der-Waerden symbols with the help of the Pauli
matrices $\sigma_{\mu}$ and ${\widetilde{\sigma}}_{\mu}\doteq\sigma^{\mu}$:
\begin{align}
{\sigma_{\mu}}^{A\Xdot}&\doteq\frac{1}{\sqrt{2}}(\sigma_{\mu})^{AX}, &
{\sigma_{\mu}}_{A\Xdot}&\doteq\frac{1}{\sqrt{2}}
({{\widetilde{\sigma}}_{\mu}}^{T})_{AX}.
\label{def:vdw}
\end{align}
They satisfy the following relations
\begin{align}
{\sigma_{\mu}}^{A\Xdot}{\sigma_{\nu}}_{A\Xdot}
&=\eta_{\mu\nu} &
{\sigma_{\mu}}_{A\Xdot}{\sigma^{\mu}}_{B\Ydot}
&=\epsilon_{AB}\epsilon_{\Xdot\Ydot}
\label{eq:vdw}
\end{align}
With the help of these we can build the infinitesimal spinor
transformations
\begin{equation}
g\approx\eins+\frac{1}{2}\theta_{\alpha\beta}S^{\alpha\beta},
\label{eq:spininf}
\end{equation}
with the generators
\begin{equation}
{(S^{\alpha\beta})^{A}}_{B}
={\sigma^{[\alpha}}^{A\Xdot}{\sigma^{\beta]}}_{B\Xdot}.
\label{eq:spingen}
\end{equation}
Note that the $ \sigma $'s are hermitian: $
\overline{{\sigma_{\mu}}^{A\Xdot}}={\sigma_{\mu}}^{X\dot{A}} \newcommand{\Bdot}{\dot{B}} $. Again
we define the projectors for the tensor product. But we have only two
irreducible parts:
\begin{align}
{{P_S}^{AB}}_{CD}&=\frac{1}{2}
(\delta^A_C\delta^B_D+\delta^A_D\delta^B_C), &
{{P_\epsilon}^{AB}}_{CD}&=\frac{1}{2}
\epsilon^{AB}\epsilon_{CD}, \\
P^2=P, &
P_S+P_\epsilon&=\eins.
\end{align}
We get the following identities:
\begin{align}
S^{\alpha\beta}S_{\alpha\beta}
&=\overline{S}^{\alpha\beta}\overline{S}_{\alpha\beta}
=-3\eins, \label{eq:sgen1} \\
S^{\alpha\beta}\otimes S_{\alpha\beta}
&=\overline{S}^{\alpha\beta}\otimes\overline{S}_{\alpha\beta}
=4P_\epsilon-\eins \label{eq:sgen2}, \\
S^{\alpha\beta}\otimes\overline{S}_{\alpha\beta}
&=\overline{S}^{\alpha\beta}\otimes S_{\alpha\beta}
=0. \label{eq:sgen3}
\end{align}
In order to have (\ref{def:a}) in a pure spinor representation we have
to decompose the tensor $\xbar$ into spinor indices according to
\begin{equation}
x^{A\Xdot}\doteq x^\mu{\sigma_{\mu}}^{A\Xdot}.
\end{equation}
Assume $t\widetilde{x}$ transforms under the $u$-fold tensor product of
\ensuremath{\mathit{SL}(2,\menge{C})}\ times the $v$-fold tensor product of $\overline{\ensuremath{\mathit{SL}(2,\menge{C})}}$ then,
for infinitesimal transformations, (\ref{def:a}) yields:
\begin{equation}
B^{\alpha\beta}
=(S^{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\eins\otimes\dots\otimes\overline{S}^{\alpha\beta})a,
\end{equation}
with $B^{\alpha\beta}$ from equation \eqref{eq:atensor} in the
corresponding spinor representation. The sum consists of $u$ summands
with one $S^{\alpha\beta}$ and $v$ summands with one
$\overline{S}^{\alpha\beta}$ with $u,v>n$. Multiplying again with the
generator and contracting the indices gives twice the Casimir on the
r.h.s. Inserting (\ref{eq:sgen1}-\ref{eq:sgen3}) yields:
\begin{multline}
(S_{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\eins\otimes\dots\otimes\overline{S}_{\alpha\beta})B^{\alpha\beta} \\
=\biggl(
-3(u+v)\eins
+2\sum_{1\leq i<j\leq u}(4P_{\epsilon_{ij}}-\eins) +
2\sum_{1\leq i<j\leq v}(4P_{\overline{\epsilon}_{ij}}-\eins)
\biggr)a.
\end{multline}
The sum over $u$ runs over $\frac{u}{2}(u-1)$ possibilities and
similar for $v$, so we find the induction:
\begin{multline}
a=\frac{1}{u(u+2)+v(v+2)}\Biggl[
-(S_{\alpha\beta}\otimes\dots\otimes\eins+\dots
+\eins\otimes\dots\otimes\overline{S}_{\alpha\beta})B^{\alpha\beta}+
\\
8\biggl(
\sum_{1\leq i<j\leq u}P_{\epsilon_{ij}}
+\sum_{1\leq i<j\leq v}P_{\overline{\epsilon}_{ij}}
\biggr)a \Biggr].
\label{eq:spinind}
\end{multline}
It already contains the induction start for $a^{(AB)}, a^{(XY)}$ and
$a^{A\Xdot}$.
\section{General covariant BPHZ subtraction}
\label{sec:bphz}
In this section we shrink the distribution space to $\Scal'$ since we are
dealing with Fourier transformation. Let $x,q,p\in\Rvm$. BPHZ subtraction
at momentum $q$ corresponds to using $w=e^{iq\cdot}$ in the
EG subtraction \cite{pap:prange1}.
\begin{align}
\widehat{\tR{\omega}{e^{iq\cdot}}}(p)
&\doteq\scp{\tR{\omega}{e^{iq\cdot}}} {e^{ip\cdot}} \\
&=\scp{^{0}t}{e^{ip\cdot}-\sum_{|\alpha|\leq\omega}
\frac{(p-q)^\alpha}{\alpha!}\ensuremath{\partial}^q_{\alpha}e^{iq\cdot}}.
\label{eq:BPHZ}
\end{align}
It is normalized at the subtraction point $q$, i.e.:
$\ensuremath{\partial}_{\alpha}\widehat{\tR{\omega}{e^{iq\cdot}}}(q) = 0,\ |\alpha| \leq
\omega$. This is always possible for $q$ totally spacelike,
$(\sum_{j\in I}q_{j})^2<0, \forall I\subset\{1,\dots,m\}$
\cite{pap:ep-gl,priv:duetsch}. In massive theories one can put $q=0$
and has the usual subtraction at zero momentum which preserves
covariance. But this leads to infrared divergencies in the massless
case. There we can use the results from above to construct a covariant
BPHZ subtraction for momentum $q$ by adding
$\sum_{|\alpha|\leq\omega} \frac{i^{|\alpha|}}{\alpha!} a^\alpha
p_\alpha$ to (\ref{eq:BPHZ}), according to equation \eqref{def:tRGcov}.
For $|\beta|\geq\omega+1$, ${^0t}x^\beta$ is a well defined distribution
on $\Scal$ and so is $\ensuremath{\partial}_\beta\widehat{^0t}$.
\subsection{Lorentz invariance on \Rv}
We have
\begin{align}
V_k e^{iq\cdot}&=e^{iq\cdot}\sum_{m=0}^k\frac{1}{m!}(-iqx)^m, &
\ensuremath{\partial}_\sigma V_k e^{iq\cdot} &=iq_\sigma e^{iq\cdot}\frac{1}{k!}(-iqx)^k.
\end{align}
Inserting this into (\ref{eq:amun}) we find:
\begin{multline}
a^{(\mu _1\dots\mu_n)}=\frac{i^n(-)^{\omega+1}}{(\omega-n)!}
\frac{ (n-1)!!}{(n+2)!!}
q_{\sigma_1}\dots q_{\sigma_{\omega-n}}
\sum_{s=0}^{\left[\frac{n-1}{2} \right] }
\frac{(n-2s)!!}{(n-2s-1)!!}
\left(q_\rho\ensuremath{\partial}^\rho\ensuremath{\partial}^{(\mu_1}-q^{(\mu_1}\square\right)\times \\
\times \eta^{\mu_2\mu_3}\dots\eta^{\mu_{2s}\mu_{2s+1}}
\ensuremath{\partial}^{\mu_{2s+2}}\dots\ensuremath{\partial}^{\mu_n)}\square^s
\ensuremath{\partial}^{\sigma_1}\dots\ensuremath{\partial}^{\sigma_{\omega-n}}\widehat{{^0t}}(q).
\label{eq:amunq}
\end{multline}
\subsubsection*{Example}
Take the setting sun in massless scalar field theory:
$^{0}t=\frac{i^{3}}{6}D_{F}^{3} \Rightarrow \omega=2$.
\begin{align*}
a^\mu&=-\frac{i}{3}(q_\sigma q_\rho \ensuremath{\partial}^\rho \ensuremath{\partial}^\sigma \ensuremath{\partial}^\mu
-q^\mu q_\sigma\ensuremath{\partial}^\sigma\square)\widehat{^{0}t}(q), \\
a^{\mu\nu}&=\frac{1}{4}(q_\rho \ensuremath{\partial}^\rho \ensuremath{\partial}^\mu \ensuremath{\partial}^\nu
-q^{(\mu}\ensuremath{\partial}^{\nu)}\square)\widehat{^{0}t}(q),
\end{align*}
and adding $ip_{\mu}a^{\mu} - \frac{1}{2}p_{\mu}p_{\nu}a^{\mu\nu}$
restores Lorentz invariance of the setting sun graph subtracted at $q$.
\subsection{General induction}
We only have to evaluate $B^{\alpha\beta}$ with $w=e^{iq\cdot}$ and
plug the result into the induction formulas \eqref{eq:spinind} and
\eqref{eq:atenrec}.
\begin{equation}
B^{\alpha\beta}
=2i^{n}(-)^{\omega+1}\sum_{j=1}^{m}\sum_{|\gamma|=\omega-n}
\frac{q^{\gamma}}{\gamma!}q_{j}^{[\alpha}\ensuremath{\partial}_{j}^{\beta]}
\ensuremath{\partial}_{\gamma}\widetilde{\ensuremath{\partial}}\,\widehat{^{0}t}(q).
\end{equation}
Here, $q_{j}$ are the $m$ components of $q$ hence $\gamma$ is a $4m$
index and $\alpha,\beta$ are four indices. The tensor (spinor)
structure of $\widetilde{\ensuremath{\partial}}$ is given by $\widetilde{x}$ in
\eqref{eq:atensor}.
\section{Summary and Outlook}
The subtraction procedure in EG renormalization makes use of an
auxiliary (test) function and hence breaks Lorentz covariance, since
no Lorentz invariant test function exists. But this symmetry can be
restored by an appropriate choice of counterterms. We give an explicit
formula for their calculation in lowest order and an inductive one for
higher orders. Using the close relationship to BPHZ subtraction this
directly translates into a covariant subtraction at totally spacelike
momentum.
We expect our solution to be useful for all calculations for which the
central solution ($w=1$, see \cite{bk:scharf}) does not exist, namely
all theories that contain loops of only massless particles.
\section{Acknowledgment}
I would like to thank Klaus Bresser and Gudrun Pinter for our short
but efficient teamwork and K. Fredenhagen for permanent support.
\clearpage
|
\section{Introduction}
With the advent of HERA there has been a great increase in the scope
of the theoretical effort to understand the physics of hadronic
scattering at high energy.
This is a challenging subject especially as it might provide a
bridge between perturbative partonic physics of short distance
processes (e.g. DIS at moderate $x_{Bj}$) and soft physics of hadronic
states which, presumably, dominate the high energy asymptotics.
The borderline between the two --- the "semihard" physics --- is
interesting also in its own right.
In essence it is the physics of partonic systems which are, on one hand
dense enough for new collective phenomena to play important role but,
on the other hand are perturbative since the average momentum transfer
between the partons is high enough.
In this semihard regime one expects to see perturbativelly
controllable nonlinear effects which depart from the standard linear
evolution of DGLAP \cite{dglap} or BFKL \cite{bfkl} type and
subsequently lead to unitarization of hadronic cross sections.
An approach to this high partonic density regime from the partonic
side has been spearheaded by Levin and collaborators \cite{levin}
based on earlier work by Mueller \cite{mueller} and is an "all
twist" generalization of the GLR recombination picture \cite{glr,mq}.
It lead to the formulation of a nonlinear evolution equation which
exhibits a perturbative mechanism of unitarization.
Analysis of this equation suggests that already at present HERA
energies the nonlinearities in the gluon sector are considerable and
linear evolution for gluons should break down.
Better experimental data on gluon distributions would be extremely
valuable in order to verify/falsify this assessment.
A complementary approach was pioneered some years ago by McLerran and
Venugopalan \cite{mv}.
It was later somewhat reshaped conceptually and considerably developed
technically in a series of papers \cite{jkmw,jklw,moap,soap}.
Here the idea is that in the high density regime rather than using
partonic language, it is more appropriate to use the language of
classical fields.
The hadron then is considered as an ensemble of configurations of the
gluon field.
The statistical weight that governs the contributions of different
configurations to the ensemble averaging changes when one probes the
hadron on different time scales.
Decreasing $x_{Bj}$ corresponds to increasing the time resolution and
therefore corresponds to probing the hadron on shorter time scales.
The change of the statistical weight with $x_{Bj}$ is governed by the
evolution equation.
As long as the field intensity is large enough this evolution should
be perturbative in $\alpha_s$ but essentially nonlinear in the field
intensity itself.
This evolution equation to first order in $\alpha_s$ was derived in
\cite{jklw,moap}.
We will refer to it as the JKLW equation in the following.
In \cite{soap} the double logarithmic limit of this evolution was
considered.
It was shown that in this limit the evolution of the gluon density
becomes unitary at large density.
Qualitatively, the evolution is very similar to that discussed in
\cite{levin}, although the details are different.
A detailed numerical study of the doubly logarithmic limit of the JKLW
evolution was recently performed in \cite{jw}.
Technically the derivation of \cite{moap} is fairly involved.
Several consistency checks were performed in \cite{jklw} and
\cite{soap} to make sure that the known results are recovered from the
general evolution equation in the weak field limit.
This includes the BFKL equation, the doubly logarithmic approximation
to the DGLAP equation and the GLR equation.
It is, however, desirable to have some additional independent checks on
the equation which do not involve the weak field limit.
It is the aim of the present paper to provide one such check.
In a nutshell the issue we address is the following.
The evolution equation in \cite{jklw,moap,soap} was derived invoking a
two step procedure.
Rather than considering directly the evolution of the correlators of
the gluon field, one first considers the evolution of the colour
charge density.
In the second step, one re-expresses the evolution equation for the
charge density as the evolution equation for the vector potential
(chromoelectric field).
The evolution of the field correlators is in fact what one is after,
since it is the vector potential and not the charge density that
couples directly to fermions and, therefore, are more directly related
to physical observables.
Our observation in this paper is that one can avoid the introduction
of the colour charge density altogether and derive the evolution
equations directly for the field correlators.
This procedure has the advantage of being somewhat simpler both
technically and conceptually.
Nevertheless, the final result for the evolution should be the same as
in the two step procedure of \cite{jkmw,jklw,moap,soap}.
The comparison of our results with the earlier derived formulae
provides us with a consistency check on the calculation.
We find, in fact, that the results presented in \cite{moap} are not
entirely correct.
However, after correcting some algebraic mistakes in \cite{moap} we
show explicitly that the two approaches yield identical results.
We provide the corrected expressions for the "kernels" of the
evolution equation, which are somewhat simpler than the expressions
found in \cite{moap}.
We also clarify the issue of possible Gribov ambiguity and show
explicitly that the divergent Jacobians, which appeared in the
intermediate steps of the derivations in \cite{soap}, cancel completely
in the final expressions for the correlators of the chromoelectric
field.
Therefore, the Gribov ambiguity, although affecting the relation
between the colour charge density and the chromoectric field, does not
affect the evolution of the field correlators, at least to order
$\alpha_s$.
Since the procedure discussed in the present paper avoids the
introduction of the charge density entirely, the whole approach is
free from the Gribov problem.
Perhaps somewhat surprisingly the corrections to the results of
\cite{moap} that we find do not affect either the weak field limit
discussed in \cite{jklw} or the doubly logarithmic limit of
\cite{soap}.
They therefore have no bearing on the derivation of the BFKL equation
in our approach and also do not help to reconcile the doubly
logarithmic limit of the JKLW equation \cite{soap} with the nonlinear
equation studied in \cite{levin}.
The plan of the paper is the following. In Section 2 we briefly recap
the procedure of the derivation of the evolution equation as described
in \cite{jklw,moap,soap} and reformulate it directly in terms of the
gluon field correlators.
In Section 3, using some of the results of \cite{moap}, we calculate
the real and virtual parts of the evolution in terms of the field
correlators and provide the corrections to the results of \cite{moap}.
Finally, in Section 4 we discuss our results.
\section{The JKLW evolution}
First, let us briefly recall the framework and the results of
\cite{jklw, moap,soap}.
In this approach the averages of gluonic observables in a hadron are
calculated via the following path integral
\begin{multline}
<O(A)> = \int D\rho DA^\mu O(A)
\exp\Big\{ -\int d^2 x_\perp F[\rho ^a(x_\perp)]
-i\int d^4 x \frac{1}{4}{\rm tr} F^{\mu\nu}F_{\mu\nu} \\
+{\frac{1}{N_c}} \int d^2 x_\perp dx^- \delta (x^-)
\rho_{a}(x_\perp) {\rm tr}T_a W_{-\infty,\infty}
[A^-](x^-,x_\perp)\Big\}
\label{action}
\end{multline}
where the gluon field strenght tensor is given by
\begin{equation}
F^{\mu\nu}_{a} =
{\partial}^{\mu} A^{\nu}_{a} - {\partial}^{\nu} A^{\mu}_{a} -
g f_{abc}A^{\mu}_{b}A^{\nu}_{c}
\end{equation}
and $W$ is the Wilson line in the adjoint representation along the
$x^+$ axis
\begin{equation}
W_{-\infty ,+\infty}[A^{-}](x^- , x_{\perp}) =
P\exp\Bigg[+i g \int dx^+ A^-_a (x^+,x^- , x_{\perp}) T_a \Bigg]
\end{equation}
The hadron is represented by an ensemble of colour charges localized
in the plane $x^-=0$ with the (integrated across $x^-$) colour charge
density $\rho(x_\perp)$. The statistical weight of a configuration
$\rho(x_\perp)$ is
\begin{equation}
Z=\exp \{-F[\rho]\}
\label{z}
\end{equation}
In the tree level approximation (in the light cone gauge $A^+=0$) the
chromoelectric field is determined by the colour charge density
through the equations
\begin{equation}
F^{+i}={1\over g}\delta(x^-)\alpha^{i}(x_\perp)
\label{chrom}
\end{equation}
and the two dimensional vector potential $\alpha^i(x_\perp)$ is "pure
gauge", related to the colour charge density by
\begin{align}
{\partial}^{i}{\alpha}_{a}^{j} &-
{\partial}^{j}{\alpha}_{a}^{i} -
f_{abc}{\alpha}_{b}^{i}{\alpha}_{c}^{j}=0 \nonumber \\
{\partial}^{i}{\alpha}_{a}^{i} &
=-{\rho}_{a}
\label{sol}
\end{align}
Integrating out the high longitudinal momentum modes of the vector
potential generates the renormalization group equation, which has the
form of the evolution equation for the statistical weight $Z$
\cite{jklw}
\footnote{All the functions in the rest of this paper
depend only on the transverse coordinates. For simplicity of notation
we drop the subscript $\perp$ in the following.}
\begin{equation}
{d\over d\zeta}Z= \alpha_s \left\{{1\over 2}{\delta^2
\over\delta\rho(u)\delta\rho(v)}\left[Z\chi(u, v) \right] -
{\delta\over\delta\rho(u)}\left[Z\sigma(u)\right]\right\}
\label{final}
\end{equation}
In the compact notation used in Eq.~(\ref{final}), both $u$ and $v$
stand for pairs of colour index and transverse coordinates, with
summation and integration over repeated occurrences implied. The
evolution in this equation is with respect to the rapidity $\zeta$,
related to the Feynman $x$ by
\begin{equation}
\zeta=\ln 1/x
\end{equation}
Technically it arises as a variation of $Z$ with the cutoff imposed on
the longitudinal momentum of the fields $A^\mu$.
The quantities $\chi[\rho]$ and $\sigma[\rho]$ have the meaning of the
mean fluctuation and the average value of the extra charge density
induced by the high longitudinal momentum modes of $A^\mu$.
They are functionals of the external charge density $\rho$.
The explicit expressions have been given in \cite{moap} and it is our
aim in this paper to provide a check on these expressions.
Eq.~(\ref{final}) can be written directly as an evolution equation for
the correlators of the charge density.
Multiplying Eq.~(\ref{final}) by $\rho(x_1)...\rho(x_n)$ and
integrating over $\rho$ yields
\begin{multline}
{d\over d\zeta}<\rho(x_1)...\rho(x_n)>= \\
= \alpha_s \Bigg[\sum_{0<m<k<n+1}
<\rho(x_1)...\rho(x_{m-1})\rho(x_{m+1})...
\rho(x_{k-1})\rho(x_{k+1})...\rho(x_n)\chi(x_m, x_k)> \\
+\sum_{0<l<n+1}<\rho(x_1)...\rho(x_{l-1})\rho(x_{l+1})
...\rho(x_n)\sigma(x_l)>\Bigg]
\label{correl}
\end{multline}
This set of equations for the correlators of the colour charge density
completely specifies the evolution of the hadronic ensemble as one
moves to higher energies (or lower values of $x$).
The evolution equations for the correlators of the charge density can
be rewritten as equations for the correlators of the vector potential
\cite{soap}.
\begin{multline}
\label{correlf}
{d\over d\zeta}
<\alpha_{a_1}^{i_1}(x_1)...\alpha_{a_n}^{i_n}(x_n)> = \\
= \alpha_s \Bigg[\sum_{0<l<n+1}
<\alpha_{a_1}^{i_1}(x_1)...
\alpha_{a_{l-1}}^{i_{l-1}}(x_{l-1})
\alpha_{a_{l+1}}^{i_{l+1}}(x_{l+1})...
\alpha_{a_n}^{i_n}(x_n)\sigma_{a_l}^{i_l}(x_l)> \\
+ \sum_{0<m<k<n+1}<\alpha_{a_1}^{i_1}(x_1)...
\alpha_{a_{m-1}}^{i_{m-1}}
(x_{m-1})\alpha_{a_{m+1}}^{i_{m+1}}(x_{m+1})... \\
\times
\alpha_{a_{k-1}}^{i_{k-1}}(x_{k-1})
\alpha_{a_{k+1}}^{i_{k+1}}(x_{k+1})...
\alpha_{a_n}^{i_n}(x_n)\chi_{a_ma_k}^{i_mi_k}(x_m, x_k)>
\Bigg]
\end{multline}
The quantities $\chi_{ab}^{ij}$ and $\sigma_a^i$ have a very simple
physical meaning.
The high momentum modes of the vector field which have been integrated
out in order to arrive at the evolution equation induce extra colour
charge density $\delta\rho$.
The average value of this induced density and its mean fluctuation
appear in the evolution equations eq.(\ref{correl}) as $\sigma_a^{ }$
and $\chi_{ab}^{ }$.
The appearance of the induced colour charge density leads to the
change in the value of the chromoelectric field through the solution
of Eq.~(\ref{sol}) with $\rho+\delta\rho$ on the right hand side.
The quantities $\sigma_a^i$ and $\chi_{ab}^{ij}$
are the average value
and the mean fluctuation of the induced field respectively.
It is perhaps helpful to explain how $\sigma_a^i$ and
$\chi_{ab}^{ij}$
were obtained
in
\cite{soap}.
As shown in \cite{moap}, the induced charge density can be decomposed
into two pieces
\footnote{The reason for the notation $\tilde\rho$ rather than simply
$\rho$ will be explained in the next section.}
\begin{equation}
\delta\rho=\delta\tilde\rho_1+\delta\tilde\rho_2
\end{equation}
The first piece $\delta\tilde\rho_1$ is order $g$ while the second
piece $\delta\tilde\rho_2$ is order $g^2$.
The $\delta\tilde\rho_1$ is time dependent, and
has zero average value, while its mean fluctuation is order $g^2$.
The $\delta\tilde\rho_2$ being $O(g^2)$ contributes only to the
average value of $\delta\tilde\rho$ and not to the mean fluctuation.
Assuming that the classical equations eq.(\ref{sol}) hold not only for
the background field but also for the relevant part of the fluctuation
field one can solve those equations perturbatively.
Writing
\begin{equation}
\delta\alpha^i=\delta\alpha_1^i+\delta\alpha_2^i
\end{equation}
with $\delta\alpha_1$ being $O(g)$ and $\delta\alpha_2$ being $O(g^2)$
and keeping in the classical equations all terms to order $g^2$ we
have
\begin{align}
&D_{ab}^i\delta\alpha_{1b}^j
-D_{ab}^j\delta\alpha_{1b}^i
+D_{ab}^i\delta\alpha_{2b}^j
-D_{ab}^j\delta\alpha_{2b}^i
-f_{abc}\delta\alpha_{1b}^i\delta\alpha_{1c}^j=0\nonumber \\
&\partial^i\delta\alpha_{1a}^i+\partial^i\delta\alpha_{2a}^i
=-(\delta\tilde\rho_{1a}+\delta\tilde\rho_{2a})
\label{solp}
\end{align}
We have defined for convenience
\begin{align}
\alpha_{ab}^i &= f_{abc}\alpha^i_c \nonumber \\
D_{ab}^i &= \partial^i\delta_{ab}+\alpha^i_{ab}
\label{def}
\end{align}
To order $g$ we find
\begin{equation}
\delta\alpha^i_1=-D^i{1\over \partial D}\delta \tilde\rho_1
\label{corresp1}
\end{equation}
Therefore, to order $g^2$
\begin{equation}
\chi_{ab}^{ij}(x,y)=
r_{ac}^i(x,u)\chi_{cd}^{ }(u,v)r_{db}^{\dagger j}(v,y)
\label{indcor}
\end{equation}
with
\begin{equation}
r_{ab}^i(x,y)=
- <x|[D^i\frac{1}{\partial D}]_{ab}|y>
\label{r}
\end{equation}
Here $<x|O|y>$ denotes a configuration space matrix element in the
usual sense.
At order $g^2$ we have
\begin{equation}
\delta\alpha^i_2=-D^i{1\over \partial D}
\delta \tilde\rho_2-{1\over 2}\epsilon^{ij}\partial^j
{1\over D\partial}\delta\alpha_{1}\times\delta\alpha_{1}
\label{corresp2}
\end{equation}
Here the cross product is defined as
\begin{equation*}
A\times B = f_{abc}\epsilon^{ij}A_a^iB_b^j
\end{equation*}
We thus have
\begin{equation}
\sigma_{a}^{i}(x)=
r_{ab}^i(x,u)\sigma_{b}^{ }(u)+p_{abc}^i(x,u,v)\chi_{bc}^{ }(u,v)
\label{indfield}
\end{equation}
with
\begin{equation}
p^i_{abc}(x,y,z)=
- \frac{1}{2}(\epsilon^{ij}\partial^j[\frac{1}{D\partial}]_{ad})(x,u)
f_{dfe}\epsilon^{kl} r^k_{fb}(u,y)r^l_{ec}(u,z)
\label{p}
\end{equation}
The procedure of deriving eq.(\ref{correlf}) employed in
\cite{moap,soap} consists, therefore, of two steps.
One first splits the gluon field into the classical background field
$\alpha^\mu$ and the fluctuation field $a^\mu$.
The modes of the fluctuation field with longitudinal momenta in some
range $\alpha_s\ln {\Lambda^+\over \Lambda^{'+}}$ are assumed to be
small.
One defines operatorialy the induced charge density $\delta\rho$ in
terms of the fluctuation fields $a^\mu$ and the quantities $\sigma$
and $\chi$ are calculated by integrating out the fluctuation fields
perturbatively.
In the second step, one solves classical equations of motion which
include
the induced charge density and calculates $\sigma^i$ and $\chi^{ij}$.
Clearly, consistency requires that the two step procedure that leads
from eq.(\ref{action}) through eqs.(\ref{final},\ref{correl}) to the
evolution equations eq.(\ref{correlf}, \ref{indcor}, \ref{indfield})
must be equivalent to the following.
Start with the equivalent of eq.(\ref{action})
\footnote{We note that the statistical weight $Z[\alpha_i]$ which
appears in eq.(\ref{actionf}) is not equal to $Z$ of eq.(\ref{z})
since going from eq.(\ref{action}) to eq.(\ref{actionf}) involves the
change of variables $\rho\rightarrow\alpha_i$. The two statistical
weights, therefore, differ by an appropriate Jacobian.}
\begin{multline}
<O(A)> =\int D\alpha^i DA^\mu O(A) Z[\alpha_i(x_\perp)]
\exp\{ -i\int d^4 x {1\over 4}{\rm tr} F^{\mu\nu}F_{\mu\nu} \\
-{{1}\over{N_c}} \int d^2 x_\perp dx^- \delta (x^-)
\partial^i\alpha_a^i(x_\perp) {\rm tr}T_a W_{-\infty,\infty}
[A^-](x^-,x_\perp)\}
\label{actionf}
\end{multline}
Integrate out the high longitudinal momentum components of $a^\mu$ as
before, but instead of calculating the induced charge density
$\sigma_a^{ }$ and $\chi_{ab}^{ }$, calculate directly the induced
chromoelectric field $\sigma_a^i$ and $\chi_{ab}^{ij}$.
Technically this calculation is somewhat simpler, since there is no
need to consider the operator $\delta\rho$, which is nonlinear in the
fluctuation field $a^\mu$.
Instead, one directly calculates the distribution of the static
component of $a^\mu$.
The resulting evolution equations should coincide with
eq.(\ref{correlf}).
With this formulation one circumvents completely the need to introduce
the colour charge density $\rho$ and to solve classical equations
for $\alpha^i$ in terms of $\rho$.
While one may want to introduce $\rho$ for reasons of convenience, our
present understanding is that it is not necessary from the point of
view of physics.
The physics that our approach is meant to address
is that of the evolution of the hadronic ensemble.
The relation between $\alpha^i$ and $\rho$ on the other hand is
supposed to hold at every value of $\zeta$, and therefore itself is
unrelated to evolution in $\zeta$.
The concept of $\rho$ may be sometimes useful to formulate models for
the statistical weight $Z$ at some particular value of $\zeta$ as was
the original motivation of \cite{mv}.
This could then serve as an initial condition for the evolution.
This, however, is a separate question and we do not intend to address it
here.
Before we proceed further, we wish to make one more comment about the
relation between the chromoelectric field and the colour charge
density eqs.(\ref{corresp1},{\ref{corresp2}).
Both these equations contain the dangerous factor $(\partial
D)^{-1}$.
The operator $\partial D$ has zero, as well as negative eigenvalues
and is very reminiscent of the operators usually associated with
the Gribov ambiguity in nonabelian gauge theories.
In fact, it is quite clear that it has precisely the same origin.
The second equation in eq.(\ref{solp}) has the form of the Lorentz
like gauge fixing condition on the fluctuation field $\delta\alpha$.
Since the calculation is performed in a nonvanishing background field,
the Lorentz gauge indeed suffers from Gribov ambiguity precisely due
to negative eigenvalues of the operator $\partial D$.
Given this, one may worry that our perturbative calculation is plagued
with the Gribov ambiguity
\footnote{In standard perturbation theory, the Gribov ambiguity does
not show up in any finite order. This is due to the fact that one
expands the operator $\partial D$ and its inverse in powers of the
coupling constant. To leading order then the operator does not have
any negative eigenvalues, which ensures that no problems arise in
finite order perturbative calculations. Our situation is, however,
different. Since our background field is not assumed to be $O(g)$, the
operator cannot be expanded. Therefore, there is no guarantee that the
problem does not show up even in perturbation theory.}.
However this is not necessarily the case.
The point is that $\delta\rho$ itself is not arbitrary.
It is calculated through the fluctuation field and, at the end of the
day, is averaged over with some statistical weight
$Z[\delta\rho]$.
It could well be that the statistical weight is such that it only
allows induced charge density of the form $\delta\rho=\partial D X$
with regular $X$.
If that is the case, the dangerous denominator cancels and the induced
field is well defined and regular.
In fact, in our present formulation where the calculation is performed
directly in terms of the field, it is almost clear that this should
indeed happen.
In this setup one calculates directly $\delta\alpha$, and
eqs.(\ref{corresp1},\ref{corresp2}) should be read from right to left,
as equations determining an auxilliary quantity $\delta\rho$ through
$\delta\alpha$ rather than the other way round.
The operator $\partial D$ then appears in the numerator and all
expressions are regular.
In fact we will show in the next section by explicit calculation that
all ``dangerous factors'' indeed cancel in the final expressions for
$\chi_{ab}^{ij}$ and $\sigma_a^i$.
Note that, if one insists on formulating the problem in terms of the
colour charge density, the absence of the Gribov ambiguity implies a
nontrivial consistency condition on the statistical weight $Z[\rho]$.
Taking an arbitrary weight $Z$ will render the calculation of
chromoelectric field correlators ill defined especially at strong
fields (strong coupling). This was indeed observed in the numerical
calculation \cite{gv} where a simple Gaussian in $\rho$ was used as the
weight function
\footnote{This problem does not arise in the more recent numerical work
\cite{kv} since in effect this work uses a different definition of
$\rho$ for which the relations analogous to
eqs.(\ref{corresp1},\ref{corresp2}) do not involve singular factors.}.
In the next section we will calculate $\sigma_a^i$ and
$\chi_{ab}^{ij}$ induced by high longitudinal momentum modes.
\section{The Induced Chromoelectric Field}
The main ingredients needed for the calculation of the induced
chromoelectric field are the eigenfunctions of the quadratic action
for the small fluctuations in the static background $\alpha^i$.
Solving the classical equations of motion that follow from the action eq.(\ref{actionf}) at fixed $\alpha^i$ we find
(in the gauge $\partial^i A^i(x^+\rightarrow -\infty)=0$) the classical solution
\begin{equation}
A_{cl}^{-}=0, \quad A^i_{cl} =\alpha^i (x_\perp)\theta(x^-)
\end{equation}
Defining the quantum fluctuation field $a^\mu$ by
$A^\mu =A^\mu_{cl}+a^\mu$ and expanding the action to
second order in $a^\mu$ we have
\begin{equation}
S=\frac{1}{2g^2} \Bigg\{
a^-_{x} K_{xy} a^-_{y} +
2a^-(\partial^+ Da - 2 fa ) +
2\partial^+a^i\partial^-a^i +
a^i \bigg[ D^2 \delta^{ij} - D^{i}D^{j} \bigg] a^j
\Bigg\}
\label{eq:action}
\end{equation}
Here we are using the notation
\begin{align}
&[fa]_{a}(x^+,x^-,x_\perp ) =
\delta (x^-) \alpha^i_{ab}(x_\perp )
a^i_{b}(x^+,x^-,x_\perp ) \nonumber \\
&Da =
D^{i}[\alpha]a^{i}=(\partial^i\delta_{ab}+
\theta(x^-)\alpha^i_{ab})a_b^i
\label{cov}
\end{align}
and as previously
\begin{equation}
\alpha^i_{ab}=f_{abc}\alpha^i_c
\end{equation}
The operator $K$ is
\begin{equation}
K_{ab}^{xy} =
- \bigg[(\partial^{+})^2 \delta_{ab}
+\partial^i\alpha^i_{ab} \delta (x^-) \frac{1}{\partial^-} \bigg]
\label{K}
\end{equation}
Note that there is no ambiguity in the definition of the operator
$1/\partial^- $ in this expression. It is defined in the sense of
principal value.
This follows directly from the fact that the matrix $\alpha^i_{ab}$ is
antisymmetric and therefore the term involving $1/\partial ^-$ in
eq.(\ref{eq:action}) vanishes for zero frequency fields.
This eigenfunctions of the quadratic fluctuation operator
have been found
in \cite{moap} and we cite here
the relevant results.
The calculation is performed in the lightcone gauge $A^+=0$ with the
residual gauge freedom fixed by the condition
\begin{equation}
\partial^i A^i(x^-\rightarrow -\infty)=0
\end{equation}
It is convenient to define an auxiliary field
\begin{equation}
\tilde a^- = a^-+ K^{-1} (\partial^+ Da - 2 fa )
\label{tildea}
\end{equation}
This field can be seen to decouple from $a_i$. Its correlator is
\begin{equation}
<\tilde a^-_x\tilde a^-_y>=K^{-1}_{x,y}
\end{equation}
The operator $K$ Eq.(\ref{K}) has zero modes.
Defining the projector matrices
$\eta$ and $\mu$ by
\begin{equation}
\mu_{ab} \partial^i\alpha^i_{bc} \frac{1}{\partial^-} = 0,\quad
\eta_{ab} \partial^i\alpha^i_{bc} \frac{1}{\partial^-} =
\rho_{ac} \frac{1}{\partial^-}
\end{equation}
and
\begin{equation}
\mu + \eta = 1 ,\quad
\mu^{2} = \mu ,\quad
\eta^{2} = \eta
\label{eq:project}
\end{equation}
we can write the normalizable zero modes of $K$ in the form
\begin{equation}
f_a(x_\perp ,x^-,p^-)=\mu_{ab}f(x_\perp ,p^-)
\label{zero}
\end{equation}
The operator $K$ is therefore, strictly speaking, non invertible.
The operator $K^{-1}$ in eq.(\ref{tildea}) has to be understood as the
inverse of $K$ on the space of functions which does not include the
functions Eq.(\ref{zero}).
Further, it is only the nonzero mode part of $a^-$ that enters the
definition of $\tilde a^-$ in eq.(\ref{tildea}).
For our calculation we will need the properly normalized solutions of
the equations of motion that follow from the action
eq.(\ref{eq:action}).
The complete set of these solutions was found in \cite{moap}
\begin{multline}
a^{i}_{p^-,r} =
ge^{ip^{-}x^{+}} \int d^{2}p_{\perp} \bigg[ \theta (-x^-)
\exp\left( i{p^{2}_{\perp} \over 2p^{-}}x^{-} -
ip_{\perp}x_{\perp}\right) v^{i}_{-, r}(p_{\perp}) \\
+ \theta (x^{-}) U (x_\perp )
\exp\left( i{p^{2}_{\perp} \over 2p^{-}}x^{-} -
ip_{\perp}x_{\perp}\right)
\left[U^{\dagger}v^{i}_{+,r}\right](p_\perp )
+ \theta (x^{-}) \gamma^{i}_{+,r}\bigg]
\label{solut}
\end{multline}
The frequency $p^-$ is a good quantum number since the background
field is static.
Here $r$ is the degeneracy label, which labels independent solutions
with the frequency $p^-$.
In the free case it is conventionally chosen as the transverse
momentum, $\{r\}=\{p^i\}$.
The matrix $U(x_\perp )$ is the $SU(N)$ matrix that parametrizes the
two dimensional ``pure gauge'' vector potential $\alpha^i(x_\perp)$
\begin{equation}
\alpha^i(x_\perp )=
{i\over g}U(x_\perp )\partial^i
U^\dagger(x_\perp )\nonumber
\end{equation}
The auxiliary functions $\gamma^i_+, v^i_\pm$ are all determined in
terms of one vector function.
Choosing this independent function as $v^i_-$ we have
\begin{align}
v^{i}_{+,r}&= \bigg[T^{ij} -L^{ij} \bigg]
\bigg[t^{jk} -l^{jk}\bigg] v^{k}_{-,r} \\
\gamma^{i}_{+,r}&= 2 D^{i} \bigg[{D^j\over D^2} -
{\partial^j\over \partial^{2}}\bigg]
\bigg[t^{jk} -l^{jk}\bigg]v^k_{-,r}
\end{align}
where we have defined the projection operators
\begin{alignat}{2}
T^{ij} & \equiv \delta^{ij} -
{D^i D^j \over D^2}, &
\qquad L^{ij} & \equiv
{D^i D^j \over D^2} \nonumber\\
t^{ij} & \equiv \delta^{ij} -
{\partial^i \partial^j \over \partial^2}, &
\quad l^{ij} & \equiv
{\partial^i \partial^j \over \partial^2}
\label{eq:tproj}
\end{alignat}
The proper normalization of the eigenfunctions requires $v^i_-$ to be
chosen as complete set of eigenfunctions of the two dimensional
Hermitian operator $O^{-1}$
\begin{multline}
\qquad [(t-l)O^{-1}(t-l)]_{ab}^{ij}(x_\perp,y_\perp)= \\
<x_\perp|\delta_{ab}^{ij} -
2 \bigg[[\partial^i{1\over\partial^2}-D^i{1\over D^2}]
S^{-1}[{1\over\partial^2}\partial^j-{1\over D^2}D^j]
\bigg]_{ab}|y_\perp> \qquad
\end{multline}
such that
\begin{equation}
\int d^2 r_\perp v^{i}_{-,r,a}(x_\perp )
v^{\ast j}_{-,r,b}(y_\perp ) =
{1\over 4\pi |p^-|}[O^{-1}]_{ab}^{ij} (x_\perp ,y_\perp )
\label{ort}
\end{equation}
The rotational scalar operator $S$ is
\begin{equation}
S = {1\over D^2}+2 [{\partial^i\over\partial^2}-
{D^i\over D^2}][{\partial^i\over\partial^2}-
{D^i\over D^2}]=
{1\over D^2}-2{1\over \partial^2}
\partial\alpha{1\over D^2} +
2{1\over D^2}D\alpha{1\over\partial^2}
\end{equation}
For further use we also need the expression for the $a^-$ component of
the fluctuation field. Using the explicit expression for the operator $K$
from \cite{moap} we get from eq.(\ref{tildea})
\begin{multline}
a^-(x^-,x_\perp,p^-) =
\tilde a^- -\theta(x^-) \int_{x^-}^\infty
dy^-D^i(a^i-\gamma^i_+) \\
-\theta(-x^-)\bigg[\int_0^\infty dy^-D^i(a^i-\gamma^i_+)
+\int_{x^-}^0 dy^-\partial^ia^i\bigg]
+ 2ip^-\eta[{D^i\over D^2} - {\partial^i\over \partial^2}]
(t-l)^{ij}v^j_-(x_\perp)
\label{a-}
\end{multline}
We note that this expression differs by a $x^-$-independent constant
from the one given in
\cite{moap}. The reason is that in \cite{moap} a constant have been
subtracted from
$a^-$ such that $\int_{-\infty}^{+\infty}dx^-a^-(x^-)=0$. This
corresponds to the symmetric definition of the integral in
eq.(\ref{a-}). This is incorrect, since it violates the residual gauge
fixing
$\partial^ia^i(x^-\rightarrow -\infty)$ at the one loop level.
We will see this explicitly later in this section. At any rate,
straightforward albeit somewhat tedious calculation gives the result
eq.(\ref{a-}) and this is the expression that will be used in the rest
of this paper.
So far the formulae presented in this section
(except for the corrected expression for $a^-$ eq.(\ref{a-}))
are identical to those that appear in \cite{moap}
with the only difference that the background charge
density $\rho$ has been substituted by the background
field via $\rho=-\partial^i\alpha^i$.
Now however we will take a different route.
Our aim is to calculate the order $O(\alpha_s\ln 1/x)$ correction
to the background chromoelectric field eq.(\ref{chrom}) directly, rather than
to the background charge density.
According to the discussion in the previous section (see also
\cite{moap}), we are therefore interested in the following two
quantities
\begin{align}
\alpha_s\ln 1/x\: \chi_{ab}^{ij}(x_\perp,y_\perp)
&= <a_{a}^i(x_\perp,x^-\rightarrow\infty,
x^+)a_b^j(y_\perp,y^-\rightarrow\infty, x^+)> \\
\alpha_s\ln 1/x\: \sigma_a^i(x_\perp)
&= <a_a^i(x_\perp,x^-\rightarrow\infty,x^+)>
\end{align}
It should be noted that, since the background is static, none of the
quantities defined above depend on $x^+$.
\subsection{The real part - the mean fluctuation}
It is a straightforward matter to calculate $\chi^{ij}_{ab}$.
Recall that we need this quantity to order $g^2$. The fluctuation
fields $a^\mu$ are formally of order $g$ themselves, and therefore to
calculate the mean fluctuation we do not have to include loop
corrections.
Examining the expression for the general solution eq.(\ref{solut}) we
see that it contains oscillating pieces, which do not contribute
to the value of the field at infinity as well as the $\gamma_+$ piece,
which does not vanish at infinity and, therefore, determines the
distribution of the vector potential there.
\begin{equation}
\chi_{ab}^{ij}(x_\perp,y_\perp)=4\pi\int d p^-
<\gamma^{i}_{+,a}(x_\perp,p^-)\gamma^{j}_{+,b}(y_\perp,-p^-)>
\end{equation}
Using the explicit expressions for $\gamma_+^i$ we find after some
trivial algebra
\begin{equation}
\chi_{ab}^{ij}(x_\perp,y_\perp)=
2<x_\perp|\{{D^i\over D^2}[D^2-S^{-1}]{D^j\over D^2}\}_{ab}|y_\perp>
\label{chif}
\end{equation}
We now want to compare this with the corresponding result of \cite{moap}.
The induced charge density $\delta\rho$ in \cite{moap} is
\begin{equation}
\delta\rho=\delta\rho_1+\delta\rho_2
\end{equation}
with
\begin{multline}
\label{rho11}
\delta\rho_{1a}(x_\perp ) = f_{abc} \alpha_{b}^{i}(x_\perp )
\Bigg[a_{c}^{i}(x^-=0)-\int_0^{\infty}dx^-
\partial^+ a^i_c(x^-)\Bigg] \\
- {{1}\over{2}} f_{abc} \partial^i \alpha^{i}_{b}(x_\perp ) \int dy^+
\Bigg[\theta (y^+ - x^+) - \theta (x^+ - y^+) \Bigg]
a^{-}_{c}(y^+,x_\perp ,x^-=0)
\end{multline}
and
\begin{multline}
\label{rho21}
\delta \rho_{2a}(x_\perp ) =
f_{abc} \int d x^- [\partial^+ a_{b}^{i}(x) ]a_{c}^{i}(x) \\
- {1\over{2}} \partial^i \alpha^{i}_{b}(x_\perp )
\int\! dy^+ a^{-}_{c}(y^+,x_\perp ,x^-=0)
\int\! dz^+ a^{-}_{d}(z^+,x_\perp ,x^-=0) \\
\times \Bigg[f_{ace}f_{bde}\theta (z^+ -x^+)\theta (x^+ -y^+) +
f_{abe} f_{cde}\theta (x^+ -z^+) \theta (z^+ -y^+) \Bigg]
\end{multline}
Only $\delta\rho_1$ contributes to $\chi$.
Substituting the expressions for $a^i$ and $a^-$ into eq.(\ref{rho11})
we find
\begin{equation}
\delta\rho_1=
-2(\partial D)
[{D\over D^2}-{\partial\over\partial^2}]
(t-l)v_-
\end{equation}
Thus, we obviously have
\begin{equation}
\gamma_+^i=-D^i{1\over \partial D}\delta\rho_1
\label{relat}
\end{equation}
This reproduces exactly eq.(\ref{corresp1}).
Obviously the relation between $\chi_{ab}^{ }$ and $\chi_{ab}^{ij}$,
eq.(\ref{indcor}) is also reproduced by this result.
We note that our result for $\chi_{ab}^{ }$ is somewhat different than the
one presented in \cite{moap}. As discussed before this is due to an
incorrect treatment of the $x^-$-independent component of $a^-$ in
\cite{moap}.
\subsection{The virtual part - the average value of the field}
We now proceed to calculate the virtual part of the evolution kernel.
For this purpose we have to calculate the zero frequency part of the
$\{ij\}$ and $\{i-\}$ components of the fluctuation propagator.
The calculation of the $\{ij\}$ at zero frequency is straightforward.
The result is
\begin{multline}
\lim_{p^-\rightarrow 0}
G_{ab}^{ij}(x^-,y^-;x_\perp,y_\perp,p^-)
\equiv \lim_{p^-\rightarrow 0} <a_a^i(x^-,x_\perp,p^-)
a_b^j(y^-,y_\perp,p^-)>= \\
-i\delta^{ij} \delta(x^--y^-)
\Big [\theta(x^-)
<x_\perp|({1\over D^2})_{ab}|y_\perp>+
\theta(-x^-)<x_\perp|({1\over \partial^2})_{ab}|y_\perp>
\Big]
\end{multline}
The $\{i-\}$ component is then calculated immediately using this
result, eq.(\ref{a-}) and the fact noted earlier that the field
$\tilde a$ decouples from $a^i$.
The result is
\begin{multline}
\lim_{p^-\rightarrow 0}
G_{ab}^{i-}(x^-,y^-;x_\perp,y_\perp,p^-)
\equiv \lim_{p^-\rightarrow 0} <a_a^i(x^-,x_\perp,p^-)
a^-_b(y^-,y_\perp,p^-)>= \\
i\theta(x^--y^-)
\Big[\theta(x^-)
<x_\perp|({D^i\over D^2})_{ab}|y_\perp>+
\theta(-x^-)<x_\perp|({\partial^i\over
\partial^2})_{ab}|y_\perp>
\Big]
\end{multline}
\begin{fmffile}{moap1pics}
\unitlength=1mm
\gdef\T#1#2#3#4{
\begin{fmfgraph*}(40,16)
\fmfpen{.6thin}
\fmfi{wiggly}{(0,.5h) -- (.6w,.5h)}
\fmfi{wiggly}{fullcircle scaled .4w shifted (.8w,.5h)}
\def\V##1##2##3{
\fmfiv{dec.shape=circle, dec.size=3, lab.angle=##3,
label=\noexpand\texttt{\noexpand\small ##1}}{##2}}
\V{#1}{(0,.5h)}{95}
\V{#2}{(.6w,.5h)}{140}
\V{#3}{(.6w,.5h)}{40}
\V{#4}{(.6w,.5h)}{-40}
\end{fmfgraph*}}
\begin{figure}[t]
\centering
\mbox{ \subfigure[]{\T{$i$}{$j$}{$k$}{$l$}}\qquad
\subfigure[]{\T{$i$}{$-$}{$j$}{$j$}}\qquad
\subfigure[]{\T{$i$}{$-$}{$-$}{$-$}}}
\caption{\small One loop tadpole diagrams contributing to
$\sigma^i_a$. The tadpole is calculated at $x^-\rightarrow\infty$.}
\end{figure}
\end{fmffile}
We are now ready to calculate $\sigma_a^i$.
It is given by the one loop tadpole diagrams of Fig. 1.
The vertex 1c comes from the expansion of the Wilson line
term in the action to third order in the fluctuation.
The separate contributions of the diagrams can be written in terms of
the fluctuation propagator $G^{\mu\nu}\equiv<a^\mu a^\nu>$ in the
following form
\begin{align}
\text{1a} &= \frac{i}{2}\int d y^-d^2 y_\perp
G^{ij}_{ab}(x^-, y^-, x_\perp,y_\perp, p^-=0)
\epsilon^{jk}D^k_{bc}f_{cde} \epsilon^{mn}
G^{mn}_{de}(y^-,y^-;y_\perp,y_\perp,y^+,y^+)
\nonumber \\
\text{1b} &= - i \int d y^-d^2 y_\perp
G^{i-}_{ab}(x^-, y^-, x_\perp,y_\perp, p^-=0)
f_{bcd}{\partial}^{+}_{\tilde{y}^- = y^-}
G^{jj}_{cd}(y^-,\tilde{y}^-;y_\perp,y_\perp,y^+,y^+)
\nonumber \\
\begin{split}
\text{1c} &= \frac{i}{N_c} \int d y^-d^2 y_\perp d y^+ d w^+ d z^+ \\
& \qquad\qquad\times
\delta (y^-) ({\partial}^{i}{\alpha}^{i}_{b}(y_\perp))
G^{i-}_{ac}(x^+, y^+,x^-, y^-, x_\perp,y_\perp)
G^{--}_{de}(w^+, z^+,y^-, y^-, y_\perp,y_\perp)
\\
& \qquad\qquad\quad\times
\bigg[
\theta (z^+ - y^+)\theta (y^+ - w^+) f_{bef}f_{cdf} -
\theta (y^+ - z^+)\theta (z^+ - w^+) f_{bcf}f_{def}
\bigg]
\end{split}
\end{align}
The diagram Fig.~1a corresponds directly to the second term in
eq.(\ref{indfield}).
For this diagram we immediately find
\begin{equation}
\delta\sigma^i_{a(1)}(x)=
-{1\over 2}\epsilon^{ij}
\left[{D^j\over D^2}\right]_{ab}\negthickspace (x,y)
f_{bcd}\epsilon^{kl}\chi^{kl}_{cd}(y,y)
\label{finals1}
\end{equation}
The diagramms Fig.~1b and 1c correspond to
the first term in eq.(\ref{indfield}) and can be written as
\begin{equation}
\delta\sigma^i_{a(2)}(x)=
-{D^i\over D^2}<\delta\rho_2>
\label{stum}
\end{equation}
with $\delta\rho_2$ (cf eq.(\ref{rho21})):
\begin{multline}
<\delta\rho_2>_a =
f_{abc} \int d x^-
< (\partial^+ a_{b}^{i}(x) )a_{c}^{i}(x)>
+ \, {1\over{2}} \,
(f_{ace}f_{bde} - \frac{1}{4} f_{abe}f_{cde}) \,
\partial^i \alpha^i_b (x_\perp ) \\
\times
\int\! \frac{d\lambda}{\lambda +i\epsilon}
dp^- \frac{1}{(p^-)^2}
<a^{-}_{c}(p^-, x_\perp ,x^-=0)
a^{-}_{d}(-p^-, x_\perp ,x^-=0)>
\label{finals}
\end{multline}
Using the results for the equal time propagators from \cite{moap} we
obtain
\begin{equation}
\begin{split}
<\delta\rho_2>_a
&=
- {1\over 2}\, (f_{ace}f_{bde} - \frac{1}{4} f_{abe}f_{cde})\,
\partial^i \alpha^i_b (x_\perp) \\
&\qquad\qquad\times <x_\perp| {1\over \partial^2}+
{1\over 2}\mu{1\over D^2}\mu
-2\left [{1\over \partial^2}\alpha D+{\mu\over 2} \right ]
{1\over D^2}S^{-1}{1\over D^2}
\left [D\alpha{1\over \partial^2}
+{\mu\over 2}\right ]|x_\perp>_{cd} \\
&\quad+f_{abc}<x_\perp|
\left [t^{ij}-l^{ij}
-2\alpha^i\partial^j{1\over \partial^2} \right]
\left [\delta^{jk}-2({\partial^j\over\partial^2}
-{D^j\over D^2})S^{-1} ({\partial^k\over\partial^2}
-{D^k\over D^2})\right ] \\
&\hspace{74mm}
\times \left [T^{ki}-L^{ki}
-2{1\over \partial^2}\partial^k\alpha^i\right ]|x_\perp>_{bc}\\
&\quad +R^a(x_\perp)
\end{split}
\label{well}
\end{equation}
with
\begin{equation}
\begin{split}
R^a(x_\perp)
&=
f_{abc}\int d^2y_\perp d^2z_\perp
{d^2p_\perp d^2k_\perp\over (2\pi)^4}
{p_\perp^2\over p_\perp^2-k_\perp^2}
e^{ip_\perp (x_\perp-y_\perp)-ik_\perp (x_\perp-z_\perp)}\\
&\qquad\times \Big \{
<y_\perp|\left [t^{ij}-l^{ij}\right]
\left [\delta^{jk}
-2({\partial^j\over\partial^2}-{D^j\over D^2})S^{-1}
({\partial^k\over\partial^2}-{D^k\over D^2})\right ]
\left [t^{ki}-l^{ki}\right ]|z_\perp> \\
&\hspace{20mm}
-U(x_\perp)<y_\perp|U^\dagger\left [T^{ij}-L^{ij}\right]
\left [\delta^{jk}
-2({\partial^j\over\partial^2}-{D^j\over D^2})S^{-1}
({\partial^k\over\partial^2}-{D^k\over D^2})\right ] \\
&\hspace{85mm}\times
\left [T^{ki}-L^{ki}\right ]U|z_\perp>
U^\dagger(x_\perp) \Big \}_{bc}
\end{split}
\label{R}
\end{equation}
Here the singularity in the integrand at $p_\perp^2=k_\perp^2$ has to be
understood in the sense of the principal value
$${1\over p_\perp^2-k_\perp^2}={p_\perp^2-k_\perp^2
\over(p_\perp^2-k_\perp^2)^2+\epsilon^2}.$$
Our final result for the induced field is
given by the sum of eq.(\ref{finals1}) and eq.(\ref{stum})
(supplemented by eqs.(\ref{well},\ref{R})).
\section{Conclusions}
To summarize, the final results of this paper are eqs.(\ref{chif}) and
(\ref{stum},\ref{well}).
They supercede the corresponding results of \cite{moap} and
\cite{soap}.
We now want to comment on this result. The first thing to observe is
that the dangerous
denominator $\partial D$ does not appear in these expressions.
The Gribov problem mentioned earlier
therefore does not affect our calculation, at least
to order $\alpha_s$.
The result for the induced field differs from the corresponding formulae
in \cite{moap} and \cite{soap} in two ways.
One
reason is the improved
treatment of $a^-$ relative to \cite{moap}.
Now we are in the position to understand why the expression for $a^-$
used in \cite{moap} is inconsistent with the residual gauge fixing.
In the previous section we have calculated the induced vector
potential
far at infinity
$x^-\rightarrow\infty$. It is not much more difficult to calculate
it everywhere in space. Diagrammatically it is given by
the same diagrams as Fig.1 except the coordinate on the free end of
the propagator is some finite $x^-$. The difference in the analytic
expressions eq.(\ref{stum}) is that the surface charge density
$\delta\rho_2$ is substituted by the local charge density integrated
up to the longitudinal coordinte $x^-$
\begin{equation}
-\theta(x^-){D^i\over D^2}\int_{-\infty}^{x^-}dy^-<\delta j^+_2(y^-)>+
-\theta(-x^-)
{\partial^i\over \partial^2}\int_{-\infty}^{x^-}dy^-<\delta j^+_2(y^-)>
\label{stump}
\end{equation}
This expression makes it explicit that the induced field vanishes at
$x^-\rightarrow-\infty$.
Therefore, our calculation clearly preserves the residual gauge
condition $\partial^ia^i(x^-\rightarrow-\infty)=0$.
However if we were to subtract the zero momentum piece from the field
$a^-$ as done in \cite{moap}, the integration limits in eq.(\ref{a-})
would become symmetric $\int_x^\infty \rightarrow {1\over
2}(\int_x^\infty+\int_x^{-\infty})$.
The effect of this would be that $G^{i-}(x^-,y^-)$ would not vanish at
$x^-\rightarrow -\infty$.
It is then obvious that we would have $\partial^ia^i(x^-\rightarrow
-\infty)\ne 0$.
The expression obtained in the present paper does not suffer from this
problem.
It is consistent with the perturbative $i\epsilon$ prescription for
regulating the $1/p^+$ gauge pole used in the earlier work
\cite{kovchegov}.
Another difference between our present result and \cite{moap} is the
appearance of $D^2$ rather than $\partial D$ and $D\partial$ in the
denominators in eqs.(\ref{finals1},\ref{stum}).
This deserves an explanation.
This is also related to another point we want to address.
Comparing eq.(\ref{stum}) with eq.(\ref{indfield}) one could wonder
whether the present method of calculation of $\chi^{ij}$ is consistent
with the two step procedure of \cite{moap, soap}.
It may look like the relation between the induced field and the
induced charge density we obtained here (eq.(\ref{stum})) is different
from the equation eq.(\ref{indfield}) which was used in the previous
work.
This however is not the case.
The reason is that the $O(g^2)$ induced charge density
$\delta\tilde\rho_2$ which appears in eq.(\ref{indfield}) is not quite
the same as $<\delta\rho_2>$ in eq.(\ref{stum}).
The $\delta\tilde\rho_2$ was defined as complete $O(g^2)$ contribution
to the average of induced density.
In other words
\begin{equation}
\delta\tilde\rho_2=<\delta\rho_1+\delta\rho_2>
\label{tilnottil}
\end{equation}
with $\delta\rho_{1,2}$ defined in eqs.(\ref{rho11},\ref{rho21}).
As we discussed above, the fluctuating part of the operator
$\delta\rho_1$ is of $O(g)$ and therefore indeed $\delta\tilde\rho_1$
can be identified with $\delta\rho_1$.
However, the vacuum average of $\delta\rho_1$ is $O(g^2)$ and does
contribute in eq.(\ref{tilnottil}).
It can be shown that
\begin{equation}
<\delta\rho_{1a}>=f_{abc}\alpha_b^i<a^i_c(x^+\rightarrow\infty)>
\label{tr}
\end{equation}
This extra contribution turns $\partial^i\delta\alpha_2^i$ into
$D^i\delta\alpha_2^i$ in the
second equation in eq.(\ref{solp}) if we use $\delta\rho_2$ rather
than $\delta\tilde\rho_2$ in its right hand side.
Taking account of this we see that the procedure described in section
2 is consistent with
eqs.(\ref{chif},\ref{stum},\ref{well}).
In \cite{moap} it was assumed that $<\delta\rho_1>=0$ and thus the
extra contribution of eq.(\ref{tr}) was overlooked.
This lead to an apparent noncancellation of spurious factors ${1\over
\partial D}$ which as we see now, do indeed cancel in the final
result.
Importantly, the corrections we find vanish in the limit of weak field
considered in \cite{jklw} and also in the double logarithmic limit,
where the field is considered not necessarily weak but slowly varying
in the transverse plane \cite{soap}.
This can be seen in the following way.
Comparing eq.(\ref{a-}) to the appropriate expression in \cite{moap}
we find that the difference between the two is proportional to
$\rho$.
In the weak field limit one only needs to know $a^-$ to order $1$ and
therefore the correction is unimportant.
For slowly varying fields all terms proportional to $\rho$ are also
negligible.
Therefore, the real part - $\chi^{ij}$ - in these two limits is
insensitive to the correction we found here.
The virtual part - $\sigma^i$ - does not contribute at all in the DLA
limit.
In the weak field limit the correction is negligible since
$<\delta\rho_1>=O((\alpha^{i})^2)$ and one only needs $\delta\rho$ to
order $\alpha^i$.
\pagebreak
{\bf Acknowledgements}
We are greatful to J. Jalilian-Marian, L. McLerran and H. Weigert
for useful discussions.
The work of J.G.M. is supported by PRAXIS
XXI/BD/11277/97 grant (Subprograma
Ci\^encia e Tecnologia do 2$^{\underline o}$ Quadro
Comunit\'ario de Apoio --- Portugal).
The work of A.K. is supported by PPARC
Advanced fellowship.
|
\section{Introduction}
The question of whether short range Edwards Anderson (EA) spin glasses
share the remarkable features of the infinite range Sherrington
Kirkpatrick (SK) model \cite{SPINGLASS} is still an open one. In this
work we present Monte Carlo simulations of the $4D$ EA Ising Spin
Glass \cite{BHAYOU,REBHYO,PARRIT,CIPARI,BCPRPR,PARIRU,BERCAM} with a bimodal
distribution of the quenched couplings, performed on large lattice
volumes (thanks to the tempering and parallel tempering simulation
technique \cite{PARTEM,OPTI}), with a large number of samples, and
down to low values of the temperature $T$. In this way we are able to
obtain detailed information about the nature of the transition, to
determine with good precision critical temperature and exponents, and
to give strong evidence supporting the fact that the low-temperature
phase is mean-field-like. A great deal of effort has gone in ensuring
reliability of the data on delicate issues such as thermalization
checks and consistency of data analysis.
The paper is organized as follows: first of all we describe the model
and the parameters of our MC simulation. We then present data related
to the Binder cumulant and to the determination of $T_c$ and $\nu$.
By analyzing the overlap susceptibility we determine the value of
$\eta$. Finally we discuss in detail about the probability
distribution of the overlap $P(q)$. We present, among others,
evidence for non-triviality of single sample $P_J(q)$ and for a
non-zero value of the position of the maximum of $P(q)$, $q_{max}$, in
the thermodynamic limit.
\section{The Numerical Simulation}
It is very difficult to run reliable numerical simulation of finite
dimensional spin glasses. The main reason for such difficulties is
the presence of many meta-stable states (responsible for aging effects
as well as for many other peculiarities of spin glasses
\cite{SPINGLASS}). The Monte Carlo dynamics gets easily trapped, and
the system only probes a restricted part of phase space.
Many algorithmic solutions have been proposed to improve the speed of
thermalization of these systems. All these techniques are related to
density scaling methods (see \cite{OPTI} for a review and references):
we use here the maybe simplest implementation of these ideas, the
parallel tempering~\cite{PARTEM}, where a number of configurations of
the system are allowed to exchange their temperature (for
multi-canonical methods, that are strongly related and have in
principle an even wider range of applicability, see for example
\cite{BERG}). Thanks to parallel tempering we have been able to
thermalize systems of volume $V=10^4$ down to $T\simeq 1.2 (0.6 T_c)$.
We study the $4D$ Edwards-Anderson Ising spin glass with binary
couplings, with Hamiltonian
\begin{equation}
H \equiv - \sum_{\langle i,j \rangle} J_{ij} \sigma_i \sigma_j \ ,
\end{equation}
where the sum runs over nearest neighboring sites, the $\sigma_i$ are
$\pm 1$ Ising spins, and the couplings are quenched variables drawn
with probability $\frac12$ among the two values
$\{ -1, +1 \}$.
The {\em overlap} among two different systems is defined as
\begin{equation}
q^{\alpha, \beta} \equiv \frac{1}{V} \sum_i \sigma_i^{\alpha}
\sigma_i^{\beta} \ ,
\end{equation}
where $\alpha$ and $\beta$ denote two configurations of the system in
the same realization of the quenched disorder.
The overlap probability distribution for a given sample is
\begin{equation}
P_J(q) \equiv \mt{\delta(q-q^{\alpha, \beta})} \ ,
\end{equation}
where $\mt{...}$ denotes the usual Gibbs average.
Its average over samples is
\begin{equation}
P(q) \equiv \md{P_J(q)} \ ,
\end{equation}
and its moments are defined as
\begin{equation}
q^{(n)} = \md{\mt{q^n}} = \int dq\ q^n\ P(q) \ .
\end{equation}
We always denote by $\mt{\cdots}$ the thermal averages
and by $\md{\cdots}$ the disorder averages.
Our simulation have been performed on a set of workstations, using a
multi-spin-coding program that was inspired by the work of
\cite{RIEGER}. We have selected the parameters of our Monte Carlo and
parallel tempering runs such to guarantee a complete thermalization of
the measured observables. We will discuss this issue in some detail.
Table (\ref{T-PARAME}) summarizes the relevant parameters used in the
simulation: we give among others the number of thermalization steps,
of measurements steps, the number of different disorder realizations
and the temperature ranges investigated by tempering. Temperature
values have been chosen uniformly spaced in the interval between
$T_{min}$ and $T_{max}$.
\begin{table}
\centering
\vspace{3mm}
\begin{center}
\begin{tabular}{|c||c|c|c|c|c|c|c|} \hline
$L$ & Thermalization & Equilibrium & Samples & $N_{\beta}$ & $\delta T$ &
$T_{min}$& $T_{max}$\\ \hline \hline
3 & 100000 & 100000 & 3200 & 17 & 0.1 & 1.2 & 2.8\\ \hline
4 & 100000 & 100000 & 2944 & 17 & 0.1 & 1.2 & 2.8\\ \hline
5 & 100000 & 100000 & 1920 & 17 & 0.1 & 1.2 & 2.8\\ \hline
6 & 100000 & 100000 & 1120 & 33 & 0.05 & 1.2 & 2.8\\ \hline
8 & 100000 & 100000 & 1376 & 33 & 0.05 & 1.2 & 2.8\\ \hline
10 & 150000 & 150000 & 512 & 56 & 0.04 & 1.2 & 3.4\\ \hline
\end{tabular}
\end{center}
\caption[0]{Parameters of the Tempered Monte Carlo runs.
\protect\label{T-PARAME}}
\end{table}
We have used different methods to verify that we have
correctly thermalized the systems. Using ``parallel tempering'', one is
actually performing a generalized Markov chain where systems at
different temperatures are allowed to ``move'' in temperature-space
too. A necessary condition for the Markov chain to be effective in
de-correlating different measurements is the fact that each system
spans at least a few times all the allowed temperature range during
the simulation. In this respect we check {\em a posteriori} that the
probability of swapping temperature has been of order $0.5$ (ensuring
in this way that a single system did not get stuck at a specific value
of $T$) and that the histogram counting the time that each system has
spent at each temperature is fairly flat. This requirement is
fulfilled in all our simulations, for all $T$ and $L$ values.
Another very strong check of thermalization is the fact that the
single sample $P_J(q)$ are symmetric in the limits of the statistical
significance of the histogram. This is very well verified as can be
seen for example in figure \figref{fig:pjq} where we plot $P_J(q)$
for selected samples.
\section{The Binder Parameter, $T_c$ and $\nu$}
We start by discussing the overlap Binder parameter. We will use it to
qualify the phase transition, and to determine the critical
temperature and the first of the critical exponents, $\nu$.
We will use and describe different methods to compute the quantities
we are interested in. Our statistical sample of configurations is a
large sample, and our set of data precise (even as far as the
dependence over the lattice volume $V$ is concerned): we will show
that different analysis styles give compatible (precise) results.
We define the usual overlap Binder parameter as
\begin{equation}
g = \frac{1}{2}
\left( 3 - \frac{ \md{\mt{q^4}}}{\md{\mt{q^2}}^2}\right) \ .
\end{equation}
The Binder parameter is an adimensional quantity, and its value at the
critical point is universal. Close to $T_c$ its leading behavior is
\begin{equation}
g(L,T) \simeq \bar{g}\left(L^{\frac{1}{\nu}}
\left(T - T_c\right)\right) \ .
\label{eq:binderscaling}
\end{equation}
In usual ferromagnetic systems the infinite volume limit of the
magnetization Binder cumulant is $0$ in the warm phase (where the
distribution of the order parameter is Gaussian) and $1$ in the broken
phase: for a spin glass with replica symmetry breaking and hence a
non-trivial distribution of the overlap order parameter, the
transition is signaled by a non-trivial value of $g$ in the broken
phase (in the warm phase one expects an infinite volume limit of
zero). In both cases the location of $T_c$ is signaled by the
crossing of the curves of $g$ versus $T$ for different values of the
lattice size $L$ (asymptotically for large $L$): large $L$ curves are
lower for $T>T_c$ and higher for $T<T_c$. We show in figure
\figref{fig:binder_allsides} $g$ versus $T$ for different $L$ values.
The crossing point is close to $T\simeq 2$ for all lattice values, and
the value of the Binder cumulant at criticality, $g_c$ is close to
$0.45$. Also error analysis has been a sensitive issues. We have
always used a jackknife or a bootstrap error analysis \cite{JACKKNIFE}
{\em directly} on the fitted parameters to determine errors. Still one
has to keep in mind that statistical errors come together with
systematic errors, due to the functional form one decides to try to
fit (typically the asymptotic scaling form, that on finite size
lattices is affected by power corrections). The two types of errors
have to be kept under control separately.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{binder_allsides.eps}
\caption[qqq]{Binder parameter $g$ versus $T$, for different values of
the linear lattice size $L$ (see caption in the plot).}
\label{fig:binder_allsides}
\end{figure}
In figure \figref{fig:binder_allsides} the crossing of the different
$g_L$ curves is very clear. It is interesting to stress the
difference with the three dimensional case \cite{MAPARU} , where the
crossing at $T_c$ looks more like a merging of the different
curves. $3d$ is (very) close to the lower critical dimension, while in
$4d$ we are in a safe region: potentially this is important to make
the physical picture easier to understand.
Let us discuss a first naive approach to the data. By looking at the
crossing of the curves $g_L(T)$ versus $T$ for different $(L,L+1)$
values one sees that one cannot extract a systematic dependence of the
crossing point (and hence of the estimate of the effective critical
temperature $T_c(L,L+1)$) over $L$. Any systematic trend is smaller
than the statistical error (maybe just showing a systematic average
decrease of the estimate $T_c(L,L+1)$ when going from smaller to
larger $L$ values). The preferred value of $T_c$ is slightly larger
than $2.00$. A first naive estimate of $\nu$ can be done by
linearizing $g_L(T)$ around the estimate we have given for $T_c$, and
by evaluating the logarithm of the slope ratio (divided by the
logarithm of the two lattice sizes ratio,
$\log\left(\frac{L}{L+1}\right))$. With this method, one gets a first
estimate for a set of effective exponents $\nu(L,L+1)$. Here too, one
cannot distinguish any clear strong dependence over the lattice size:
the error one gets on $\nu$ is completely correlated to the variation
of the estimate of $T_c$. For larger $T_c$ one estimates a lower value
of $\nu$, while for lower estimates of $T_c$ one gets larger estimates
for $\nu$. The error is dominated by this effect. The estimate for
$\nu$ is close to $1$.
To get a reliable estimate of $T_c$ and of $\nu$ we have used two
methods of analysis of $g$ (see for example the discussion of the
analysis of reference \cite{INPARU,MAPARU}). In the first approach we
linearize the data close to $T_c$ (for all $L$ values) and we run a
global fit to all data: we fit $T_c$ and $\nu$ for the two variable
function $g_L(T)$ (as we said, linearized close to $T_c$). We use data
in a $T$ range around the interval $1.9-2.1$. We estimate the errors
over the fit parameters ($T_c$ and $\nu$) by a jackknife
approach~\cite{JACKKNIFE}: we repeat the fit approximately $K$ times
over a subsample of the data containing all of our statistical sample
but a fraction $\frac{1}{K}$. The error is estimated by looking at
fluctuations of the results of the $K$ fits, and by accounting for the
fact they are correlated \cite{JACKKNIFE}. We also repeat the fits by
discarding the smaller $L$ values, to check if we can observe any
systematic drift (again with good accuracy the average value of the
result does not seem to depend systematically over the $L$ range
selected). Results are very stable, and the value we estimate for
$\nu$ systematically comes out to be close to $1.10$.
In the second approach, that comes in different flavors, one only uses
data in the warm phase. This method leads to a smaller statistical
error, that is balanced from a larger systematic incertitude (since
we only select data at a given distance from $T_c$, and approaching
$T_c$ leads to a systematic drift of the estimate). In this case we
start by selecting a threshold value for $g$, $g^* \le g_c$. We start
with low values of $g^*$, and we approach $g_c$ from below: we cannot
get too close to $g_c$ or the merging of the curves for different
sizes makes the error over the measurement too large (we use values of
$g^*$ going from $=0.2$ to $0.4$. We use a polynomial fit to
interpolate the data for $g(T)$, at different $L$ values. We have
decided to use a polynomial of degrees four (we have checked it
guarantees stable fits and consistent results), and we fit a $T$ range
in the critical region (for $L=3$ we use the data in the $T$ range
$1.5-2.8$, for $L=10$ we use the range $1.88-2.16$). We define now
$T_c(L,g^*)$ as the crossing point of the fitted polynomial with the
horizontal line at $g^*$, and $\nu^*$ as
\begin{equation}
\lim_{L\to\infty} T_c(L,g^*)
= T_c(g^*) + \frac{A}{L^{\frac{1}{\nu^*}}}\ .
\end{equation}
When $g^*\to g_c$ $\nu^*\to \nu$. If $g^*$ is too small violations of
scaling are dominant, while if one approaches too much $g_c$ the
merging of the $g$ curves makes the error over the determination of
$\nu^*$ overwhelming. The errors have been estimated by using a {\em
bootstrap} approach (very similar in spirit to the jackknife
technique, see \cite{JACKKNIFE}): one emulates fake sets of data with
a Gaussian distribution around the real measurements, fits these
multiple sets of fake data and compute the errors over the fit
parameter.
We note at last that we have also used a variation of this second
method, described in \cite{MAPARU}, based on the direct analysis of
the derivative of $g$ with respect to $T$. Also this method gives
results that are compatible with the other ones.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{binderrescaled.eps}
\caption[qqq]{$g$ versus $L^{\frac{1}{\nu}}(T-T_c)$,
with $\nu=1.0$, $T_c=2.03$.}
\label{fig:binderrescaled}
\end{figure}
Our final estimates, averaged over the results obtained using these
different approaches, are
\begin{equation}
T_c = 2.03 \pm 0.03\ ,
\end{equation}
and
\begin{equation}
\nu = 1.00 \pm 0.10\ .
\end{equation}
In the rest of this paper we will use these two values as our best
estimates of $T_c$ and $\nu$. We show in figure
\figref{fig:binderrescaled} the data for $g_L(T)$ rescaled by using
these two values: the scaling turns out to be very satisfactory.
\section{The Overlap Susceptibility and $\eta$}
The determination of the overlap susceptibility, $\chi_q$, provides
various possible ways to determine the exponent $\eta$ (and hence of
the exponent $\gamma$). In a spin glass in the RSB phase the overlap
susceptibility
\begin{equation}
\chi_q \equiv V \langle q^2 \rangle
\end{equation}
is expected to diverge for all values of $T\le T_c$.
We show $\chi_q$ versus $T$ in figure \figref{fig:susc}.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{susc.eps}
\caption[qqq]{Overlap susceptibility
$\chi_q$ versus $T$, for different $L$ values.}
\label{fig:susc}
\end{figure}
The first method is
based on the fact that we expect that at $T=T_c$
\begin{equation}
\chi_q(L,T=T_c) \simeq L^{2-\eta}\ .
\end{equation}
We use a linear interpolation of the data in the region close to
$T_c$. As in the case of $g$ the error in the estimate is mainly
related to the choice of $T_c$. The fit at $T=2.03$ by using $L>3$
gives an estimate of $0.28$.
In the second method we use data where $L\gg\xi$. We go as close to
$T_c$ as possible, under the condition that data on our larger lattice
($L=10$) coincide, in our statistical accuracy, with the ones at
$L=8$. Here we expect that
\begin{equation}
\chi_q(T) \simeq (T-T_c)^{-(2-\eta)\nu}\ .
\end{equation}
We can use data down to $T=2.5$ (i.e. at a $\Delta T\simeq 0.5$ from
$T_c$), where finite size effect start to be sizable even at
$L=10$. We show our best fit (in a $T$ interval of $=.2$) in figure
\figref{fig:susc_warm}. In this region we have a stable fit, with
$\eta$ close to $-0.4$. Even if this second measurement is not very
precise (we have to stay quite far from the critical region) it is
interesting the fact that we get a coherent determination of $\eta$,
by using a completely different scaling region than in the former
analysis (the new analysis also depends on the value of $\nu$ we have
determined by using $g$).
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{susc_warm.eps}
\caption[qqq]{Best fit to the overlap susceptibility
$\chi_q$ versus $T-T_c$, at $T>T_c$.}
\label{fig:susc_warm}
\end{figure}
The last approach we use for determining $\eta$ is based on the
analysis of the scaling properties of the distribution probability
$P(q)$ of the overlap order parameter $q$ in the region $q\simeq 0$ at
$T=T_c$. We analyze the behavior of $P(q)$ in the next section, but we
discuss now the scaling of $P(0)$ of $T_c$ in order to define our
determination of $\eta$. At $T=T_c$ we expect
\begin{equation}
P(q\simeq 0) \simeq L^{\frac{d-2+\eta}{2}}\ ,
\end{equation}
i.e. in $d=4$ a scaling with $L^{\frac{2+\eta}{2}}$.
We find a very good best fit (we do not include the $L=3$ data), with
an $\eta$ value close to $-0.3$.
By considering all the methods we have discussed in this section we
give our final estimate
\begin{equation}
\eta = -0.30 \pm 0.05\ ,
\end{equation}
that we will use in the rest of our analysis. In figure
\figref{fig:suscrescaled} we plot $\chi_q$ rescaled by using our best
fits. The rescaling works fine.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{suscrescaled.eps}
\caption[qqq]{
Rescaled overlap susceptibility $\frac{\chi_{SG}}{L^{2-\eta}}$ versus
$L^{\frac{1}{\nu}}(T-T_c)$, with $T_c=2.03$, $\nu=1.0$ and
$\eta=-0.30$.}
\label{fig:suscrescaled}
\end{figure}
Let us also notice that we have a good agreement with the results
reported in \cite{PARIRU} for the $4d$ EA model with Gaussian
couplings. There the authors find $\nu\simeq 1.06$, and $\eta \simeq
-0.35$. Universality seems to work.
\section{$P(q)$}
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{pdiqlowtallsides.eps}
\caption[qqq]{$P(q)$ at $T= 1.2$ (broken phase),
for different lattice sizes.}
\label{fig:pqt12}
\end{figure}
In the two former sections we have shown that the $4D$ EA model
undergoes a phase transition, and we have determined its location and
the critical exponents. Now we will try to qualify it in better
detail, by determining and analyzing the probability distribution of
the order parameter, $P(q)$.
In figure \figref{fig:pqt12} we show our average $P(q)$ (averaged over
the different disorder realizations) at $T=1.2<T_c$. When increasing
the lattice size the peak where $P(q)$ is maximum shifts to lower $q$
values: for showing that there is a phase transition to a phase with a
non zero expectation value of $q$ we have to show that the peak does
not go to $q=0$ when $L\to\infty$. The {\em plateau} of $P(q)$ for
$q\simeq 0$ does not lower when increasing $L$, as we will discuss
better in the following. We remind the reader that in the RSB Parisi
Mean Field scenario the $P(q)$ is (in zero magnetic field) a non
trivial function, that in the infinite volume limit is formed by a
$\delta$ function at $q=q_{EA}$ and by a regular part that extends
down to $q=0$. On the contrary if the broken phase has the same
structure of the one of an ordered ferromagnet $P(q)$ has to become
asymptotically the sum of two $\delta$ functions at $\pm q_{EA}$.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{p_di_q_allT2200.eps}
\caption[qqq]{$P(q)$ at $T= 2.2$ (warm phase),
for different lattice sizes.}
\label{fig:pqt22}
\end{figure}
For sake of comparison we show in figure \figref{fig:pqt22} what
happens in the warm phase, where the average $P(q)$ shrinks to a
Gaussian distribution around $q=0$ when $L\to\infty$.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{p_di_q_alltempL8.eps}
\caption[qqq]{$P(q)$ at $L=8$,
for all values of $T$. From single peak in $q=0$ to continuous part and
double peak at large $q$ for increasing $T$.}
\label{fig:pqL8}
\end{figure}
In figure \figref{fig:pqL8} we compare $P(q)$ at different values of
$T$ on the same lattice size. From the single peaked shape at high $T$
one gets a clear double peaked structure, with a clear {\em plateau}
at low $q$, in the low $T$ region.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{qmax_vs_L1.3r0_025.eps}
\caption[qqq]{$q_{max}$ versus $L^{-1.3}$.}
\label{fig:qmL13}
\end{figure}
As we have said, in order to establish that we are having a real phase
transition in the infinite volume limit we have to show that the value
of $q=q_{max}$ where $P(q)$ is maximum does not go to zero. We start
by plotting in figure \figref{fig:qmL13} $q_{max}$ versus $L^{-1.3}$
(the exponent $1.3$ comes from our best fit, see later). It is easy to
see that an asymptotic value $q_{max}=0$ seems unplausible.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{qmax_t1300.eps}
\caption[qqq]{$q_{max}$ versus $L$ and our two best fits.}
\label{fig:qmax_t1300}
\end{figure}
Making this last statement more quantitative needs a more careful
analysis. In order to do that we fit
\begin{equation}
q_{max}(L) = q_{max}(\infty) + \frac{A}{L^{\alpha}}\ ,
\end{equation}
both with $q_{max}(\infty)=0$ and by allowing for it a non zero
value. In figure \figref{fig:qmax_t1300} we plot the values of
$q_{max}$ versus $L$ and the results of the two best fits, one with a
fitted value of $q_{max}(\infty)$ and the second with fixed
$q_{max}(\infty)=0$. This second fit is clearly unsuitable, and it has
a very high value of $\chi^2$. In the first fit we get
\begin{equation}
q_{max}(\infty) = 0.548 \pm 0.006\ ,
\end{equation}
that is our best estimate for the position of the $\delta$ function at
$q_{EA}$ in the infinite volume limit. We estimate $\alpha=1.3\pm 0.1$
(in the fit with a zero asymptotic value one finds the very small
value $\alpha\simeq 0.2$).
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{pdizeromediata1200.eps}
\caption[qqq]{$P(q\simeq 0)$ versus $L$. $T=1.2$.}
\label{fig:p0}
\end{figure}
In figure \figref{fig:p0} we show the value of $P(q)$ close to $q=0$
(averaged over a small $q$ range, where $P(q)$ is remarkably constant,
in order to diminish statistical fluctuations) as a function of
$L$. One cannot observe any statistically significant decrease of this
value for increasing large lattice volume (there is a small decrease
only for small volumes). The most plausible implication of this
evidence is that the system has many stable states, and that the cold
$T$ phase is characterized by Replica Symmetry Breaking (even if it
has to be stressed that this evidence is not as strong as the one
implied by the figure \figref{fig:qmax_t1300}).
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{2pdiq.eps}
\caption[qqq]{$P_J(q)$ for selected samples. $T=1.2$, $L=10$.}
\label{fig:pjq}
\end{figure}
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{isto_max.eps}
\caption[qqq]{
Percentage of disorder
configurations such that $P_J(q)$ has
$1$, $2$, $4$ and $6$
peaks versus $L$. The number of configurations with a complex phase
space ($P_J(q)$ with many peaks) increases strongly with $L$.}
\label{fig:isto_max}
\end{figure}
In figure \figref{fig:pjq} we plot $P_J(q)$ for selected samples, at
$T=1.2$, $L=10$. One can see here that they are very complex
distributions: such a pattern is typically related to the presence of
many states (it has to be notice however that the small side peaks are
not always there because of the presence of a real state).
To be more quantitative we have measured the percentage of disorder
configurations such that $P_J(q)$ has $1$, $2$, $4$ and $6$ peaks
versus $L$, and we plot it in figure \figref{fig:isto_max}. The
number of configurations with a complex phase space ($P_J(q)$ with
many peaks) increases strongly with $L$. We use this evidence to rule
out the picture of a {\em modified droplet model}, that has been
discussed, among others, in \cite{NS} and in references therein. The
picture of the modified droplet models implies that for each
realization of the quenched disorder there are (in the cold phase)
only two ground states, but that the value of $q_{EA}$ (i.e. the
support of the $\delta$ function that constitutes the $P_J(q)$)
depends on the sample. Here, on the contrary, the number of states for
a given sample is strongly increasing with $L$ (and with decreasing
$T$).
\section{Sum Rules}
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{lhs_rhs_sumrule.eps}
\caption[qqq]{
$\md{\mt{q^2}^2}$, $\md{\mt{q^2}}^2$
and $\md{\mt{q^4}}$ vs. $T$.
}
\label{fig:lhs_rhs_sumrule}
\end{figure}
In this section we discuss another important feature of the broken
phase of the $4D$ EA model. The starting point for this analysis can
be for example the relation:
\begin{equation}
\md{\mt{q^2}^2} =
\frac23 \md{\mt{q^2}}^2 +
\frac13 \md{\mt{q^4}} \ .
\protect\label{G-A}
\end{equation}
This is one of a set of relations that are valid in the Mean Field
Theory of Spin Glasses \cite{MPSTV}. The work contained in
\cite{SUMRULE} has established numerically that these relations are
satisfied with good accuracy also in finite dimensional spin
glasses. Following these findings a rigorous and theoretical analysis
has improved our understanding of such set of sum rules
\cite{GUERRA,AIZCON,PARISISR,NSLONG}: they are strongly related to the
ultrametric properties of the phase space.
First of all we show evidence that the relation (\ref{G-A}) has a
non-trivial content in the low-temperature phase (in the high $T$
phase it is satisfied in the form $0=0$). Figure
\figref{fig:lhs_rhs_sumrule} shows that the values of the three
quantities involved in (\ref{G-A}) are significantly different from
zero below $T_c$ (we have already shown in better detail that the
infinite volume of such quantities is non-zero).
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{lhs_minus_rhs_sumrule.eps}
\caption[qqq]{Left hand side minus right hand
side of equation \eref{G-A} vs. $T$.}
\label{fig:lhs_minus_rhs_sumrule}
\end{figure}
In figure \figref{fig:lhs_minus_rhs_sumrule} we show the difference
between the left hand side and the right hand side of \eref{G-A}. The
two contributions cancel out with good accuracy (to $3$ significant
figures), and asymptotically for large lattice size the difference
extrapolates to zero.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{ratio_sumrule.eps}
\caption[qqq]{Ratio of left hand side and right hand side
of equation \eref{G-A} versus $T$.}
\label{fig:ratio_sumrule}
\end{figure}
Another possible way to visualize the result is plotting the ratio of
the left hand side and the right hand side. As figure
\figref{fig:ratio_sumrule} shows, for $T$ below $T_c$ we get
identically one, while for $T \to \infty$ we get the value $\frac35$,
expected for a Gaussian $P(q)$.
\begin{figure}
\centering
\includegraphics[width=0.8\textwidth]{scaling_sumrule.eps}
\caption[qqq]{
Left hand side minus right hand
side of equation \eref{G-A}
versus $L$, at $T=1.4$, and
and best fit to a zero constant value with
a simple power correction.}
\label{fig:scaling_sumrule}
\end{figure}
We have also fitted the difference plotted in figure
\figref{fig:lhs_minus_rhs_sumrule}, for various values of temperatures
$T<T_c$: in all cases a fit to an asymptotic zero value with power
corrections works very well, and the exponent of the corrections is
close to $3$ for all $T$ values. As an example we plot the data
together with the best fit for $T=1.4$ in figure
\figref{fig:scaling_sumrule}.
\section{Conclusions}
In this note we have been able to give strong evidence for mean field
behavior of the $4d$ Ising spin glass with binary couplings. Life in
the $4d$ model is easier than in $3d$, where even after a large number
of intense numerical simulations the evidence for a phase transition
is still slightly marginal (even if, at this point, convincing
enough). In our case already the crossing of the Binder cumulants is
the very clear signature of a typical phase transition (as opposed to
the quasi-merging, quasi-Kosterlitz-Thouless behavior of the $3d$
case). It is clear that $3d$ is very close to the lower critical
dimension, and that there observing the effects of the physical
critical point is dramatically difficult. $4d$ is on the safer side,
and numerical simulations show that very clearly.
We have been able to determine critical exponents precisely, and
to enter in the large volume region with good accuracy. For example
we have been able to show that the peak of $P(q)$ is not going to
$q=0$ for increasing lattice size, and (with a slightly worst accuracy
and level of reliability) that the plateau at $q\simeq 0$ does not
decrease with increasing lattice size. Also we remind the reader that
non-trivial sum rules are satisfied with very good accuracy. So,
thinks look quite clear in the $4d$ case.
\section*{Acknowledgments}
We thank Giorgio Parisi, Federico Ricci-Tersenghi and
Juan Ruiz-Lorenzo for a number of interesting discussions.
|
\section{Introduction}
This is the ninth paper of a series whose purpose is to obtain Cepheid
distances to galaxies that have produced supernovae of type Ia
(SNe~Ia), thereby calibrating their absolute magnitudes at maximum
light. The Hubble diagram for SNe~Ia that are not abnormal
in either their intrinsic colors or their spectra at maximum
(\cite{bra93} 1993) is exceedingly tight (\cite{sata93} 1993;
\cite{tasa95} 1995; \cite{saha97} 1997 ; \cite{par99} 1999)
even before second-order corrections for light curve decay rate
or color (\cite{ham95} 1995, 1996a, b; \cite{rie96} 1996)
are applied. Hence, the absolute magnitude calibrations lead directly
to a good estimate
of the global value of the Hubble constant because the
SNe~Ia Hubble diagram is defined at redshifts that are well beyond any
local velocity anomalies in the Hubble flow.
The previous papers of this series concern Cepheids in IC~4182 for
SN~1937C (\cite{sanetal92} 1992, Paper~I; \cite{saha94} 1994,
Paper~II), NGC~5253 for the two SNe~Ia 1895B and 1972E
(\cite{sanetal94} 1994, Paper~III; \cite{saha95} 1995, Paper~IV),
NGC~4536 for SN~1981B (\cite{saha96a} 1996a, Paper~V); NGC~4496A for
SN~1960F (\cite{saha96b} 1996b, Paper~VI); NGC~4639 for SN~1990N
(\cite{sanetal96} 1996, Paper~VII; \cite{saha97} 1997, Paper~VIII).
The purpose of this paper is to set out the data for the discovery and
photometry of Cepheids in NGC~3627, the parent galaxy of the type Ia
SN~1989B. This case, together with the new SN~Ia 1998bu in NGC~3368
for which \cite{tan95} (1995) have a Cepheid distance, and the new Cepheid
distance of NGC~4414 (\cite{tur98} 1998), parent galaxy to SN~1974G, and
the compilation of the extant photometric data on SN~1998bu
(\cite{sun98} 1998) and on SN~1974G (\cite{schae98} 1998), now increase the
number of SNe~Ia calibrators from seven in \cite{sanetal96} (1996) and
\cite{saha97} (1997) to nine here. It will be recalled that we have used the
absolute magnitude of SN~1989B in two previous discussions
(\cite{sanetal96} 1996; \cite{saha97} 1997) but based on the then unproven
premise that the \cite{tan95} (1995) Cepheid distance to NGC~3368 would be
the same as the distance to NGC~3627, as both galaxies are in the
loose Leo association.
SN~1989B was discovered by visual inspection of NGC~3627 by
\cite{eva89} (1989) on January 30, 1989 which was 7 days before maximum light
in $B$. A detailed light curve in {\it UBVRI} and a determination of
the reddening of $E(B-V) = 0\fm37 \pm 0\fm03$ was made by \cite{wel94} (1994),
following the earlier analysis by \cite{bar90} (1990). The magnitudes at
maximum corrected for the extinction are $B({\rm
max}) = 10\fm86 \pm 0\fm13$ and $V({\rm max}) = 10\fm88 \pm 0\fm10$.
The parent galaxy NGC~3627 (M~66) is one of the brightest spirals
(Sb(s)II.2) in the complicated region of the Leo group, first isolated
by \cite{hum56} (1956, Table~XI) where 18 possible members were
identified including NGC~3627 and NGC~3368 (M~96). Both galaxies are
illustrated in the Hubble Atlas (\cite{san61} 1961, panels 12 and 23), and
the Carnegie Atlas (\cite{sanbed94} 1994, Panels 118, 137, and S14).
\cite{dev75} (1975) has divided the larger Leo group complex into three
groups. G9 in his Table~3 is a spiral--dominated subgroup near
NGC~3627 (M~66). His G11 is dominated by NGC~3368 (M~96), parent to
SN~1998bu
and contains NGC~3351, for which a Cepheid distance is also
available (\cite{gra97} 1997). The agreement of the Cepheid distances
of NGC~3368 and NGC~3351 to within $0\fm36 \pm 0\fm25$ is
satisfactory. The smaller group G49 surrounds NGC~3607. The division
into three subgroups is supported by the three-dimensional
hierarchical clustering analysis of \cite{mat78} (1978). A catalog of 52
possible members of the subgrouping in the field of NGC~3368,
NGC~3379, and NGC~3384 is given by \cite{ferg90} (1990).
Because of the complication of three subgroups rather than a well
defined single group, we have never been totally convinced that the
distance to NGC~3627, required to calibrate SN~1989B, could be taken
to be that of the Cepheids in NGC~3368 from \cite{tan95} (1995), although we
analyzed our calibration data in Papers~VII and VIII on that
premise. This assumption is now no longer necessary with the discovery
and analysis of Cepheids in NGC~3627.
Figure~\ref{fig1} shows a ground-based image of NGC~3627 made from the
Mount Wilson 100-inch Hooker blue plate (E40) that was used in the
Hubble Atlas and for one of the frames of the Carnegie Atlas. The
position of the four WFPC2 chips of the $HST$ is superposed to show
our search area for the Cepheids. The position of SN~1989B is marked,
taken from the photographs in \cite{bar90} (1990).
Figure~\ref{fig2} is a color composite $HST$ montage of stacked
frames. The dust lanes are striking both here and in Fig.~\ref{fig1}.
The extinction clearly varies over short scales on the image. This
forewarns about the importance of differential extinction to the
Cepheids, requiring good color information for them to obtain a
reliable true modulus. The extinction in NGC~3627 is more severe than
in any galaxy of our six previous calibrations in Papers~I--VIII of
the series.
The journal of the $HST$ observations and the photometry of the master
template frame are in the next section. The identification and
classification of the variables are in \S~3. The apparent
period-luminosity relations, the analysis of the severe absorption
problem and how we have corrected for it, the resulting distance
modulus, and the absolute magnitude at maximum for SN~1989B are in
\S~4. In \S~5 we combine the data with our six previous calibrators,
as well as SN~1974G in NGC~4414 and SN~1998bu in NGC~3368 from
external sources.
This paper provides a direct Cepheid distance to NGC~3627 and removes the
uncertainty, thereby strengthening the weight of SN1989B in the overall
calibration.
A first value of $H_0$ is given. A discussion of
the second-parameter corrections and their effect on $H_0$ is in the
penultimate section \S~6.
\section{Observations and Photometry}
\subsection{The Data}
Repeated images of the field in NGC~3627, shown in Fig.~\ref{fig2}, were
obtained using the WFPC2 (\cite{hol95a} 1995a) on the $HST$
between November 1997 and January 1998. There are 12
discrete epochs in the $F555W$ passband, and 5 epochs in the $F814W$
passband, spanning a period of 58 days. The duration of this period is
constrained by the time window during which this target can be
observed with $HST$ without altering the field orientation. This
window was further curtailed due to logistics for accommodating a
campaign with the NICMOS camera, for which a change of the telescope
focus was necessary, thus rendering the telescope useless for other
kinds of observations.
The epochs were spaced strategically over this period to provide
maximum leverage on detecting and finding periods of Cepheid variables
over the period range 10 to 60 days. Each epoch in each filter was
made of two sub-exposures taken back-to-back on successive orbits of
the spacecraft. This allows the removal of cosmic rays by an
anti-coincidence technique. The images from various epochs were
co-aligned to within 3--4 pixels on the scale of the PC chip, which is
1--2 pixels on the scale of the other three wide-field chips. The
journal of observations is given in Table~\ref{tbl1}.
\subsection{Photometry}
The details of processing the images, combining the sub-exposures for
each epoch while removing cosmic rays and performing the photometry
with a variant of DoPHOT (\cite{schec93} 1993) optimized for WFPC2 data has
been given in \cite{saha96a} (1996a) and need not be repeated here. The data
reduction procedure is identical to that described in Paper~V, with
the one exception of a change in the definition of the ``partial
aperture''. Instead of the $9 \times 9$ pixel aperture, a circular
aperture of 5 pixel radius was used, and the local background is
defined as the value for which the aperture growth curve is flat from
6 to 8 pixels. The details of these changes are given in \cite{ste98} (1998)
in their \S~2.2. This change produces no known systematic differences, but
improves the S/N with which aperture corrections are measured.
In keeping with the precepts in Paper~V, the measurements in any one
passband are expressed in the magnitude system defined by
\cite{hol95b} (1995b) that is native to the WFPC2. These are the $F555W$ and
$F814W$ ``ground system'' magnitudes calibrated with $HST$ ``short''
exposure frames. The issue of the discrepancy of the photometric
zero-points for the ``long'' and ``short'' WFPC2 exposures, originally
found by \cite{ste95} (1995) is described in some detail in Paper~V. In any
eventual accounting for this zero-point correction, one must {\it add}
$0\fm05$ in {\it both} passbands to the \cite{hol95b} (1995b) calibration
whenever the exposures are longer than several hundred seconds. The
cause of this zero-point difference is not fully understood at the time
of this writing, and, as in previous papers of this series, we
continue to present the basic photometry (Tables~\ref{tbl3} and
\ref{tbl4}) on the ``uncorrected'' \cite{hol95b} (1995b) ``short exposure''
calibration. And, because all of our WFPC2 observations have exposure
times that are ``long'', we make the $0\fm05$ adjustment only at the
resulting distance modulus of NGC~3627 (cf. \S~4.2.2.).
Correspondingly the distance moduli in Table~\ref{tbl5} are corrected
to the `long' calibration scale
by $0\fm05$, except the moduli of IC~4182 and NGC~5253 which were
observed with the older WF/PC.
\section{Identification and Classification of the Variable Stars}
Armed with measured magnitudes and their reported errors at all
available epochs for each star in the object list, the method
described by \cite{saha90} (1990) was used to identify variable stars.
The details specific to WFPC2 data have been given in various degrees
of detail in Papers~V, VI, and VIII.
All variable stars definitely identified are marked in
Fig.~\ref{fig3}. However, some of the identified variables cannot be
seen in Fig.~\ref{fig3} because of their extreme faintness and/or
because of the large variation in surface brightness over the field.
Hence, to complement these charts, we set out in Table~\ref{tbl2} the
X and Y pixel positions for all variable stars as they appear in the
images identified in the $HST$ data archive as U3510701R and
U3510702R.
The photometry on the \cite{hol95b} (1995b) ``short exposure''
calibration system for the final list of 83 variable stars is
presented in Table~\ref{tbl3} for each epoch and each filter. The
periods were determined with the \cite{laf65} (1965) by using only the
$F555W$ passband data. Aliassing is not a serious problem for periods
between 10 and 58 days because the observing strategy incorporated an
optimum timing scheme as before in this series.
The resulting light curves in the $F555W$ passband, together with
periods and mean magnitudes (determined by integrating the light
curves, converted to intensities, and then converting the average back
to magnitudes, and called the ``phase-weighted intensity average'' in
\cite{saha90} 1990), are shown in Fig.~\ref{fig4}, plotted in the
order of descending period.
Four objects, C2-V23, C2-V25, C2-V36, and C2-V37 are definitely
variable but are unlikely to be Cepheids. They may be periodic with
periods greater than the time spanned by our 12 epochs, or they may be
transient variables such as novae. They are plotted in
Fig.~\ref{fig4} with periods set artificially at 100 days (much larger
than the observing time base of 58 days) for the purpose of
visualization. Four obvious objects, C1-V1, C2-V38, C4-V3, and C4-V14
have periods that are just a little larger than the observing time
base. Best guesses of the periods have been made from the light curve
shapes. As the periods of these variables are not definitive, they
should be used with caution in deriving distances. The remaining
variables have light curves and periods that are consistent with being
Cepheids.
The available data for the variables in $F814W$ were folded with the
ephemerides derived above using the $F555W$ data. The results are
plotted in Fig.~\ref{fig5}. The long-period variables, transient
variables, and long-period Cepheids (as discussed above) are shown
with the same assigned ephemerides as in Fig.~\ref{fig4}. Not all the
variables discovered from the $F555W$ photometry were found in the
$F814W$ images. The fainter variables, either because they are
intrinsically faint or else appear faint due to high extinction, may
not register clearly on the $F814W$ frames which do not reach as faint
a limiting magnitude as those in $F555W$. Since photometry of such
objects was obviously impossible in $F814W$, these variables are
dropped from Fig.~\ref{fig5} and also from further analysis. Only the
68 variables from the $F555W$ frames that were recovered in at least
one of the $F814W$ epochs are considered further.
The mean magnitudes in $F814W$ (integrated as intensities over the
cycle) were obtained from the procedure of \cite{lab97} (1997) whereby
each $F814W$ magnitude at a randomly sampled phase is converted to a
mean value $\langle{F814W}\rangle$ using amplitude and phase
information from the more complete $F555W$ light curves. Note that
each available observation of $F814W$ can be used independently to
derive a mean magnitude. Hence, the scatter of the individual values
about the adopted mean $F814W$ value is an {\it external} measure of
the uncertainty in determining $\langle{F814W}\rangle$. This value is
retained as the error in $\langle{F814W}\rangle$, and propagated in
the later calculations.
The prescription given in Paper~V for assigning the light-curve
quality index QI (that ranges from 0 to 6) was used. In this scheme,
two points are given for the quality of the $F555W$ light curves, two
points for the evenness in phase coverage of the five $F814W$
observation epochs, and three points for the amplitude and phase
coherence of the $F814W$ observations compared with the $F555W$ light
curve. Hence, a quality index of 6 indicates the best possible light
curve quality. A quality index of 2 or less indicates near fatal flaws
such as apparent phase incoherence in the two passbands. This is
generally the indication that object confusion by crowding and/or
contamination by background is likely.
Table~\ref{tbl4} lists the characteristics of the 68 objects whose
light curves in $F555W$ are consistent with those of Cepheids, and for
which an $F814W$ measurement exists for at least one epoch. The
$F555W$ and $F814W$ instrumental magnitudes of Table~\ref{tbl3} have
been converted to the Johnson $V$ and Cousins (Cape) $I$ standard
photometric system by the color equations used in Papers V to VIII of
this series, as set out in equations (2) and (3) of Paper~V, based on
the transformations of \cite{hol95b} (1995b).
The magnitude scatter $\sigma_{\langle{V}\rangle}$ in Table~\ref{tbl4}
is based on the estimated measuring errors in the photometry of the
individual epochs. The determination of the scatter
$\sigma_{\langle{I}\rangle}$ is described above. The quality index
discussed above is also listed. Other columns of Table~\ref{tbl4} are
explained in the next section.
\section{The Period-Luminosity Relation and the Distance Modulus}
\subsection{The P-L Diagrams in $V$ and $I$}
As in the previous papers of this series we adopt the P-L relation in
$V$ from \cite{mad91} (1991) as
\begin{equation}
M_{V} ~~=~~ -2.76 ~\log P - 1.40~,
\end{equation}
whose companion relation in $I$ is
\begin{equation}
M_{I} ~~=~~ -3.06 ~\log P - 1.81~.
\end{equation}
The zeropoint of equations (1) and (2) is based on an adopted LMC
modulus of 18.50.
The P-L relations in $V$ and $I$ for the 68 Cepheids in
Table~\ref{tbl4} are shown in Fig.~\ref{fig6}. The filled circles
show objects with periods between 20 and 58 days that have a quality
index of 4 or higher. These are the best observed Cepheids. No
selection based on color has been made here, hence the total range of
differential extinction values is contained in the data as plotted,
explaining part of the large scatter.
The continuous line in each of the two panels shows equations (1) and
(2) as the ridge-line relation using an apparent distance modulus of
30.2. The faint dashed upper and lower envelope lines indicate the
expected scatter about the mean due to the intrinsic width of the
Cepheid instability strip in the HR diagram (\cite{sata68} 1968).
The large observed scatter of the data outside these envelope lines
are due to the combination of (1) measuring and systematic errors due
to background and contamination, (2) the random error of photon
statistics, and (3) the large effects of the variable extinction
evident from Fig.~\ref{fig2}. Any modulus inferred directly from
Fig.~\ref{fig6} would be unreliable. No matter how the lines defining
the P-L strip are shifted vertically in the two panels of
Fig.~\ref{fig6}, a large fraction of the points of the total sample
will remain outside the boundaries of the instability strip. Analysis
of the scatter is the subject of the next two subsections.
\subsection{Deriving the Distance Modulus}
\subsubsection{A Preliminary Analysis of the P-L Relation}
Inspection of Fig.~\ref{fig6} reveals that the deviations from the
ridge-lines in $V$ and $I$ of individual Cepheids are correlated.
Stars that deviate faintward in $V$ also generally deviate faintward
in $I$ and vice versa. This, of course, is a signature of variable
extinction, but it can also be caused if the {\it systematic}
measuring errors are correlated due, for example, to confusion or to
errors in the compensation for background contamination that are not
independent in the $V$ and $I$ passbands. To explore these
possibilities and to correct for them we use the tools developed in
Paper~V and used again in Papers~VII and VIII.
For each Cepheid we calculate the apparent distance moduli separately
in $V$ and in $I$ from the P-L relations of equations (1) and (2) and
the observed $\langle{V}\rangle$ and $\langle{I}\rangle$ magnitudes
from Table~\ref{tbl4}. These apparent distance moduli, called $U_{V}$
and $U_{I}$ in columns (7) and (8) of Table~\ref{tbl4}, are calculated
by
\begin{equation}
U_{V} ~~=~~ 2.76 ~\log P + 1.40 + \langle{V}\rangle~,
\end{equation}
and
\begin{equation}
U_{I} ~~=~~ 3.06 \log P + 1.81 +\langle{I}\rangle~.
\end{equation}
They are the same as equations (6) and (7) of Paper~V.
If the differences between the $V$ and $I$ moduli are due solely to
reddening, and if the dependence of the reddening curve on wavelength
is the normal standard dependence as in the Galaxy, then the true
modulus $U_T$ is given by
\begin{equation}
U_{T} ~~=~~ U_{V} - R'_{V} \cdot (U_{V} - U_{I})~,
\end{equation}
where $R$ is the ratio of total to selective absorption,
$A_{V}/E(V-I)$. This is equation (8) of Paper~V. However, equation
(5) is valid only if the difference between $U_{V}$ and $U_{I}$ is due
to extinction, not to correlated measuring errors, in which case the
value of $R$ would {\it not} be given by the normal extinction curve
where $A_{V}/A_{I} = 1.7$ and the ratio of absorption to reddening is
$R'_{V} = A_{V}/E(V-I) = 2.43$ (\cite{schef82} 1982). Such coupled
errors clearly exist (based on Fig.~\ref{fig7} later where the slope
of the $U_{V}$ vs. $U_{I}$ correlation is closer to 1 than to the
required slope of $A_{V}/A_{I} = 1.7$ if the correlation were to be
due entirely to differential extinction rather than to measuring
errors). Hence, the interpretation of the values calculated from
equations (3), (4), and (5) is considerably more complicated than
would be the case in the absence of the correlated systematic
measuring errors. Nevertheless, in the initial pass at the data to
derive the true modulus, we use equations (3) to (5) as a first
approximation. The second approximation, based on knowledge gained by
the methods of this section, is in the next section 4.2.2.
The values of $U_{T}$ are listed in column 9 of Table~\ref{tbl4}.
These would be the true moduli, as corrected for normal extinction,
assuming that there are no systematic measuring errors. The total rms
uncertainty for each $U_{T}$ value is listed in column 10. This
uncertainty includes contributions from the estimated random measuring
errors in the mean $V$ and $I$ magnitudes, (in columns 4 and 6), as
propagated through the de-reddening procedure, as well as the
uncertainty associated with the intrinsic width of the P-L relation,
i.e. a given Cepheid may not be on the mean ridge-line of the P-L
relation. The de-reddening procedure amplifies the measuring
errors. Therefore many Cepheids are needed to beat down these large
errors (notice the very large values in column 10) in any final value
of the modulus. The values shown in column 10 of Table~4 were
calculated using equations 9, 10 and 11 of Paper~V, and correspond to
$\sigma_{tot}^2$ as defined in Paper~V. However, note that equation
(11) in Paper~V should be :
\begin{displaymath}
\sigma_{width}^2 = (R'_{V}-1)^2 \cdot \rho_{V}^2 + {R'_{V}}^2 \cdot
\rho_{I}^2~.
\end{displaymath}
Our records show that while this equation was given incorrectly in
Paper~V, the calculations were done with the correct relation.
Various arithmetics done on the $U_{T}$ values in column 9 of
Table~\ref{tbl4} give the first indication of the true modulus.
Consider first all Cepheids of all quality indices, but excluding
C1-13 and C3-V15 because of their extreme values of $U_{T}$. The 66
Cepheids in this sample give $\langle{U_{T}}\rangle = 29.90 \pm 0.08$
(the error estimate is based on the adopted rms of a single Cepheid of
$0\fm646$). Restricting the sample to only those Cepheids with good
to excellent data defined by a quality index of 4 or higher gives a
subsample of 41 Cepheids for which $\langle{U_{T}}\rangle = 30.09 \pm
0.085$ (error based on an rms of $0\fm535$). Note that the 3-sigma
upper and lower limits on the true modulus from this arithmetic are
30.34 and 29.83. These are therefore very strong upper and lower
limits on the true distance modulus of NGC~3627.
These are the best values we can derive without making cuts in the
data according to period, QI, and/or color that would select the
bluest and least reddened, and that would be least affected by
selection bias at the low-period end. We consider now such
subsamples, first using cuts in period and QI, but not yet in color
which follows in the next section.
A plot of $U_{T}$ vs. period (not shown) reveals a trend that objects
with the shortest periods yield smaller $U_{T}$ moduli. This result
has been seen by us in the previously analyzed galaxies, and by other
investigators analyzing similar data in yet other galaxies. It is due
to a combination of selection bias at the short-period end
(\cite{san88} 1988) as well as non-symmetrical observational bias in
measuring colors near the faint limit, skewing $U_{T}$ via the color
effect in equation (5). We note that this trend disappears once
periods are restricted to longer than 25 days. We also note that the
variation of $U_{T}$ with Quality Index is not as acute for these data
as we have seen in previous cases. A likely reason is that $I$
magnitudes were obtained at five epochs, while in all except one of
our previous papers fewer epochs were available. At any rate there is
no compelling trend in $U_{T}$ once the sample is restricted to
objects with QI $\geq 3$.
Using the sample of 27 Cepheids that have periods $\geq 25$ days but
shorter than the baseline of 58 days, and that have QI $\geq 3$, and
weighting the individual $U_{T}$ values by $(1/{\rm rms})^2$, gives
the mean de-reddened modulus of $(m - M)_0 = 30.04 \pm 0.12$. If we
make more restrictive cuts by accepting only objects with QI $\geq 4$,
and then $\geq 5$, and then 6, (with the same period cuts as above),
we obtain respectively weighted mean ``true'' moduli of $29.99 \pm
0.13$, $30.10 \pm 0.15$, and $30.07 \pm 0.18$. This shows the general
stability of the result. For the unweighted average of the 27
Cepheids we obtain $(m - M)_0 = 30.10 \pm 0.14$. Due to the fact that
the P-L relations in $V$ and $I$ have non-negligible width at a
constant period, a Cepheid that is {\it intrinsically} redder and
fainter will, on the average, carry {\it larger} measuring
errors. This can contribute to systematically underestimating the
distance when individual $U_{T}$'s are weighted as above. The
unweighted solution is therefore preferred, since the uncertainties
are similar.
We note again that the $U_{T}$ values so derived depend on the
assumption that the differences between $U_{V}$ and $U_{I}$ are due to
reddening alone, in the absence of appreciable systematic and
correlated measuring errors, or when the errors for $U_{V} - U_{I}$
{\it are distributed symmetrically}. If equation (5) is used for
Cepheids {\it where correlated and/or asymmetrical errors in $V$ and
$I$ dominate over differential reddening}, thereby producing a ratio
of the $V$-to-$I$ errors that is different from 2.43, the $U_{T}$
derived via equation (5) {\it will be in error}.
In particular, several Cepheids which were discovered in $V$ are too
faint in $I$ to be measured, as already mentioned. This introduces a
selection effect that biases against Cepheids with bluer colors. The
effect is most pronounced at short periods where the {\it intrinsic}
colors are bluest. This effect gives an asymmetrical distribution of
errors in $U_{V} - U_{I}$ in the sense that it makes the de-reddened
modulus {\it too small}.
In Paper~V we devised a method to test for the presence of
differential extinction or for the fact that the scatter about the P-L
relation is due predominantly to measuring errors, or a combination of
both. The method, shown in Fig.~11 of Paper~V for NGC~4536 and
explained in the Appendix there, was used in Paper~VI for NGC~4496A
(Fig.~9 there) and in Paper~VIII for NGC~4639 (Fig.~7 there). The
method is to plot the difference in the apparent$V$ and $I$ moduli for
any given Cepheid as ordinate against the apparent $V$ modulus as
abscissa. If there is a systematic trend of the data along a line of
slope $dU_{V}/d(U_{V} - U_{I}) = 2.43$, then equation (5) applies and
there is clearly differential reddening. If, on the other hand, there
is a general scatter with no trend, that scatter is dominated by
measuring errors. While in the latter case true differential
extinction can be hidden by measuring errors, trying to correct for
putative reddening will result in interpreting any asymmetry in the
error distribution as specious extinction. In two of the three
previous cases, NGC~4536 (Paper~V), and NGC~4496A (Paper~VI) there is
no trend along a differential reddening line. In the third case of
NGC~4639 (Paper~VIII), there is a slight trend but also large scatter
showing that the spread of points appears to be due to a mixture of
measurement errors as well as from differential extinction.
The diagnostic diagram just described is shown for the NGC~3627 data
from Table~\ref{tbl4} in Fig.~\ref{fig7}. The filled circles show
Cepheids with periods between 25 and 58 days. The solid line indicates
the reddening vector for the P-L ridge line, if the true (de-reddened)
distance modulus is 30.05 which is close to the mean derived earlier
in this section. The dashed lines show the bounds due to the
intrinsic dispersion of the P-L relation as explained in Paper~V. The
slope of the lines is $A_{V}/E(V-I) = 2.43$ as before.
There is only marginal evidence from Fig.~\ref{fig7} for a general
trend along the solid line. The spread of points clearly spills
outside these bounds, indicating that a very significant fraction of
the scatter is due to measuring errors and related biasses. If such
errors are distributed symmetrically, equation (5) will yield the
correct answer, but if there are correlated errors in $V$ and $I$, or
if there are selection effects that depend on color, using equation
(5) will introduce errors. Note that the scatter of points is skewed
along a direction orthogonal to the reddening vector: the lower right
side appears to be more sparsely filled, indicating a possible
selection effect. A more detailed inspection of the spread in
Fig.~\ref{fig7} shows that the largest scatter occurs in the reddest
Cepheids, using the $(\langle{V}\rangle - \langle{I}\rangle)$ colors
computed from columns (3) and (5) of Table~\ref{tbl4}. This implies
that the reddest Cepheids are so more because of skewed measurement
errors than due to bona-fide reddening. This is not to deny the
presence of differential reddening, but an acknowledgement that
equation (5) alone is not adequate for obtaining a bias free
result. We proceed by making an additional restriction of the data by
color. Note that such a restriction used in conjunction with equation
(5) does not introduce a procedural bias in the distance modulus.
\subsubsection{The Distance Modulus By Restricting The Data By Color}
A plot (not shown) of the color-period relation from the data in
Table~\ref{tbl4} shows a distinctive separation into two major color
groups, one close to the intrinsic $({\langle{V}\rangle}_0 -
{\langle{I}\rangle}_0)$-period relation known for unreddened Cepheids
in the Galaxy, LMC, and SMC as summarized in \cite{sanetal99} (1999)
from data by \cite{dea78} (1978), \cite{cal85} (1985), and
\cite{fern90} (1990). The other group of Cepheids with
$(\langle{V}\rangle - \langle{I}\rangle)$ colors larger than 1.15 are
far removed from the intrinsic domain in the color-period plot. They
are also the Cepheids that show the largest deviation faintward in the
P-L relations of Fig.~\ref{fig6}. Excluding these as the Cepheids
with the largest reddening leaves a subsample of 29 Cepheids (the blue
group) with $QI \geq 3$, and with $1.15 < \log P < 1.76$. We also
have excluded C1-V13 because it is obviously an outlier. C1-V1,
although blue, is excluded because its proposed period of 75 days is
outside the baseline of 58 days.
Figure~\ref{fig8} shows again that differential reddening is not the
major factor in the scatter of the P-L relation, where $U_{V}$ is
plotted vs. $U_{I}$ for the 29 blue Cepheids of the subsample.
There is a clear correlation of $U_{V}$ and $U_{I}$, but the slope is
not $A_{V}/A_{I} = 1.7$ as required if the Cepheids below the ridge
lines of Fig.~\ref{fig6} were fainter because of a larger differential
extinction. Rather, the slope near 1 in Fig.~\ref{fig8} can only be
due to {\it correlated measuring errors} as we suspected in the last
section. Of course, a part of the correlation must also be due to
reddening, if for no other reason than the obvious dust pattern in
Fig.~\ref{fig1} and \ref{fig2}.
Figure~\ref{fig9} shows the P-L relations for the subset of the 29
bluest Cepheids. The scatter is markedly reduced from that in
Fig.~\ref{fig6}, showing that the color cut has produced a subset with
the smallest extinction and/or measuring error.
We use the $U_{V}$ and $U_{I}$ apparent moduli in Table~\ref{tbl4}
calculated from equations (3) and (4) of \S~4.2.1, and calculate mean
values $\langle{U_{V}}\rangle$ and $\langle{U_{I}}\rangle$ using the
29 Cepheids of this subset. Assuming that the measuring errors for
this sample are random and that they cancel in the mean permits the
premise that the {\it difference} in the mean apparent moduli in the
$V$ and $I$ passbands {\it is} now due to reddening. Multiplying the
mean modulus difference by $A_{V}/E(V-I) = 2.43$ then gives $A_V$,
which when subtracted from $\langle{U_{V}}\rangle$ gives the true
modulus. This, of course, is what equation (5) does automatically,
hence we need only analyze the $U_{T}$ values in Table~\ref{tbl4} for
the 29 Cepheid subsample.
The weighted mean $\langle{U_{T}}\rangle_{W}$ for the subset of 29
Cepheids gives $\langle{U_{T}}\rangle = 30.12 \pm 0.11$. The
individual $U_{T}$ values are plotted vs. $\log P$ in
Fig.~\ref{fig10}. As noted in the last section and seen in
Fig.~\ref{fig10}, there is a tendency for the shortest period Cepheids
to have the smallest individual moduli. Making the period cut at
$\log P > 1.25$ removes the tendency and gives
\begin{equation}
{\langle{U_{T}}\rangle}_W = 30.17 \pm 0.12~,
\end{equation}
as the weighted mean from the 25 Cepheids with $1.25 < \log P < 1.78$,
which we adopt. The unweighted mean is $\langle{U_{T}}\rangle = 30.24
\pm 0.09$.
Justification for our restriction to the subsample of 25 Cepheids is
given in Fig.~\ref{fig11} which is the diagnostic diagram of
Fig.~\ref{fig7} but using only this subsample. Plotted are again the
individual apparent moduli $U_{V}$ vs. the difference between the
individual $V$ and $I$ apparent moduli.
Figure~\ref{fig11} is much cleaner than Fig.~\ref{fig7}, and the data
points now scatter nearly symmetrically about the differential
reddening line.
Applying, as in previous papers of this series, the correction for the
``long'' vs. ``short'' exposure effect of $0\fm05$ to equation (6),
the de-reddened modulus of NGC~3627 is
\begin{equation}
(m - M)_0 = 30.22 \pm 0.12~,
\end{equation}
which we adopt.
\section{The Absolute Magnitude at Maximum of SN~1989B Added to
Previous Calibrators; the Present Status of the Calibration}
The light curves in the {\it UBVRI} passbands of SN~1989B are well
determined near maximum light, giving $B_{\rm max} = 12.34 \pm 0.05$,
and $V_{\rm max} = 11.99 \pm 0.05$ (\cite{wel94} 1994). These authors
have also determined the reddening of SN~1989B itself to be $E(B-V) =
0.37 \pm 0.03$. Hence, the de-reddened magnitudes at maximum light
are $B_{\rm max}^{0} = 10.86 \pm 0.13$ and $V_{\rm max}^{0}= 10.88 \pm
0.10$. The absolute magnitudes at maximum for SN~1989B are
\begin{equation}
M^0_B({\rm max}) = -19.36 \pm 0.18~,
\end{equation}
and
\begin{equation}
M^0_V({\rm max}) = -19.34 \pm 0.16~,
\end{equation}
based on equation (7) and on the Cepheid zero points of equations (1)
and (2).
These values are combined in Table~\ref{tbl5} with our previous
calibrations of the six SNe~Ia from Papers I--VII of this series. In
addition, two new values are included from data for SN~1974G in
NGC~4414 (\cite{tur98} 1998; \cite{schae98} 1998), and SN~1998bu in
NGC~3368 (M~96) whose distance modulus is by \cite{tan95} (1995) with
photometry of the SN reported by \cite{sun98} (1998). All Cepheid
distances from the WFPC2 are corrected by $0\fm05$ for the short
vs. long exposure photometric zeropoint difference (\cite{ste95} 1995;
\cite{saha96a} 1996a).
Neglecting the listed $M_B$ of SN~1895B, which is the most uncertain
of the group, and because it is not absolutely certain whether it was
spectroscopically normal at all phases, gives the straight mean values
of $\langle{M_B}\rangle = -19.49 \pm 0.03$ and $\langle{M_V}\rangle =
-19.49 \pm 0.03$. We adopt the {\it weighted} means, giving the mean
calibration without any second parameter corrections for decay rate
(see \S~7) as
\begin{equation}
\langle{M_B(\rm max)}\rangle = -19.49 \pm 0.07~,
\end{equation}
and
\begin{equation}
\langle{M_V(\rm max)}\rangle = -19.48 \pm 0.07
\end{equation}
on the Cepheid distance scale of equations (1) and (2). Inclusion of
SN~1895B would give $\langle{M_B}\rangle = -19.54 \pm 0.06$ for the
straight mean and $\langle{M_B}\rangle = -19.52 \pm 0.07$ for the
weighted mean.
We have been criticized for using calibrators such as SN~1960F and
SN~1974G, for which the photometry is more uncertain than others. We
should point out that the net worth is not just the uncertainty in the
photometry of the supernova, but the combined uncertainty with that of
the Cepheid distance determination. Given the range of uncertainties
in the Cepheid distance determinations (which depend on
distance/faintness, crowding, etc), even photometric uncertainties
considerably worse than for SN~1960F can be tolerated if the Cepheid
determinations are as good as they are for its host galaxy
NGC~4496A. In addition, we weight the contribution of individual
calibrators by the inverse variance from the combined uncertainty of
distance and supernova photometry, a procedure consistent with
Bayesian inference. Thus even bona-fide `poor' cases like SN~1974G
(which has also a relatively poor Cepheid distance to its host galaxy
NGC~4414), enter only with appropriately lowered weight.
Equations (10) and (11), based now on eight calibrators (neglecting
SN~1895B), are similar to equations (12) and (13) of Paper~VIII
(\cite{saha97} 1997) based there on six calibrators (again neglecting
SN~1895B). Equation (10) here is $0\fm03$ fainter than in Paper~VIII.
Equation (11) here is identical with that of Paper~VIII. Because of
the similarity of the equations (10) and (11) here with equations (12)
and (13) in Paper~VIII, the interim value of the Hubble constant, sans
decay-rate corrections, for this stage of our $HST$ experiment is
nearly identical with the values set out in Table~7 of Paper~VIII.
The means of all values in that table, depending on how the SNe~Ia
Hubble diagram is divided between {\it spirals} observed before and
after 1985, and/or with redshifts larger or smaller than $\log v_{220}
= 3.8$, and using equation (10) here rather than equation (12) of
Paper~VIII, are then
\begin{equation}
\langle{H_0(B)}\rangle = 58 \pm 2~~({\rm internal})\ksm~,
\end{equation}
and
\begin{equation}
\langle{H_0(V)}\rangle =59 \pm 2~~({\rm internal})\ksm~.
\end{equation}
\section{The Hubble Diagram of SNe~Ia and the Value of $H_0$
if Second Parameter Corrections are Made}
The Hubble diagram for blue SNe~Ia with $(B_{\rm max}-V_{\rm max}) <
0.20$ , based on many sources, historical as well as modern, was given
in Paper~VIII (\cite{saha97} 1997, Fig.~10). When calibrated with the
SNe absolute magnitudes, this diagram gives the Hubble constant
directly. The details of the sample are being published in a paper by
\cite{par99} (1999), and we make use of these data in this section to
calibrate again correlations of absolute magnitudes of SNe~Ia with the
often adopted second parameters of decay rate of the light curve,
color of the SNe at maximum, and galaxy type.
The precept adopted in the early papers of this series was that by
restricting both our calibrator SNe~Ia and the Hubble diagram of
SNe~Ia to ``Branch normal'' events (\cite{bra93} 1993), we have
selected a homogeneous sample of SNe~Ia for which any systematic
variation of absolute magnitude among the sample will be small enough
to be neglected to first order at the 10\% level (\cite{cad85} 1985;
\cite{lei91b} 1991b; \cite{sata93} 1993; \cite{tasa95} 1995). In the
meantime, thanks to the observational program of the Calan/Tololo
Chilean consortium (\cite{ham95} 1995; 1996a, b) and the theoretical
insights of variations in the pre-explosion conditions of the
progenitors (cf. \cite{vHipp97} 1997, \cite{hof98} 1998, \cite{nad98}
1998 for summaries), it has become clear that there is in fact a {\it
continuous} variation of SNe~Ia properties, including absolute
magnitude, that can be detected from observed second- parameter
properties and which can now be accounted for even at this $< 10\%$
level.
The most apparent of these second parameters is the change in the
shape of the light curve, quantified by the rate of decay from maximum
light (\cite{rus74} 1974; \cite{dev76} 1976; \cite{psk77} 1977, 1984;
\cite{phi87} 1987; \cite{bar90} 1990; \cite{bra92} 1992; Hamuy et al.\
1996a, b). \cite{phi93} (1993) derived a very steep relation between
decay rate and absolute magnitude using distances determined by a
variety of methods and including also very red SNe~Ia. A much flatter
dependence was found by \cite{tasa95} (1995) for blue SNe~Ia using
more reliable relative distances from recession velocities. The
flatter slope was confirmed for only the blue SNe~Ia by the subsequent
extensive Calan/Tololo data (Hamuy et al.\ 1996a, b; \cite{saha97}
1997).
A detailed discussion of the second order corrections to the SNe~Ia
distance scale depending on second and third parameter correlations is
the subject of the accompanying paper by \cite{par99} (1999) where it
is shown that besides the decay rate, color at maximum is also a
principal second parameter, confirming a similar result by
\cite{tri98} (1998).
The Parodi et al.\ paper lists relative kinematic absolute magnitudes
as if the local Hubble redshift-to-distance ratio is the same for
redshifts smaller than $10,000\kms$ as for the remote $H_0$ for $v >
10,000\kms$, neglecting the evidence set out in Paper~VIII
(\cite{saha97} 1997, \S~8; cf. also \cite{zeh98} 1998) that the local
value may be 5\% to 10\% larger than the global value.
This possible change of the Hubble ratio outward, decreasing with
distance by $\leq 10\%$ for $v$ out to $ 10,000\kms$, is also
consistent with other external data on first ranked cluster galaxies
(\cite{lau94} 1994; \cite{tam98a} 1998a). The suggestion is also
consistent with the derived shallow slope of $d\log v/dm = 0.192$ of
the local ($v < 10,000\kms$) Hubble diagram by Parodi et al.\ (1999,
their eqs. 7 and 8) rather than 0.200 required if $H_0$ did not vary
with distance.
In the analysis of this section we consider the consequences of a
variable $H_0$ decreasing outward, that is required to give a slope of
0.192 to the Hubble diagram locally, and then carry again the analysis
of the second parameter effects of decay rate, color, and galaxy type
using the derived kinematic absolute magnitudes.
\subsection{Kinematic Absolute Magnitudes Using a Variable Hubble
Constant with Local Distance for $v < 10,000\kms$ }
We accept the premise that the slope of the Hubble diagram for
redshifts smaller than $10,000\kms$ is 0.192 (\cite{par99} 1999). If
we adopt an arbitrary global (remote field) value of $H_0 = 55$ (to be
adjusted later by our calibrators in Table~\ref{tbl5}), then, by an
obvious calculation, the variation of $H_0$ in the distance interval
of $1000 < v < 10,000$ is well approximated (within $1\%$) by
\begin{equation}
H_0(v) = -5.39 \log v + 76.50~.
\end{equation}
This gives $H_0 = 60.3$ at $v = 1000\kms$ and $H_0 = 54.9$ at $v =
10,000\kms$.
Equation (14) has been used to recalculate the absolute magnitudes of
all SNe~Ia in the fiducial sample in Table~1 of \cite{par99} (1999)
for which $\log v < 4.00$. $H_o = 55$ has been assumed for $\log v >
4.00$. The results are set out in Table~\ref{tbl6} which is divided
into three parts according to Hubble type to better understand the
type dependence of the Hubble diagram seen in Fig.~10 of Paper~VIII
(\cite{saha97} 1997). The first section of Table~\ref{tbl6} lists the
SNe in late type spirals ($T$ of 3 and greater, meaning Sb to Im
types). The second section for $T = 1$ and 2 are for Sa and Sab
parent galaxies. The third section lists parent galaxy $T$ types of 0
and smaller (E and S0).
Columns (1) through (6) repeat data from Table~1 of \cite{par99}
(1999) with magnitudes at the respective maxima denoted by $B_0$,
$V_0$, and $I_0$. The redshifts in column (3) are corrected for
peculiar motions. For $v < 3000\kms$ the redshifts, reduced to the
frame of the centroid of the Local Group (\cite{yah77} 1977), were
then corrected again to the frame of the Virgo cluster using the
self-consistent Virgocentric infall model with the local infall vector
(actually retarded expansion of the Local Group relative to Virgo) of
$220\kms$ (\cite{kra86ab} 1986a, b). For $v > 3000\kms$ an additional
correction of $630\kms$ relative to the CMB frame due to the CMB
dipole anisotropy (\cite{bou81} 1981; \cite{wil84} 1984) was applied
according to the model of \cite{tasa85} (1985, their Fig.~2). For
more distant galaxies we have adopted the corrected velocities from
\cite{ham96a} (1996a). The sources for the photometry in columns (4)
to (6) are listed in \cite{par99} (1999). The magnitudes are
corrected for Galactic absorption (\cite{bur84} 1984).
Column (7) is the adopted local value of $H_0$ calculated from
equation (14) for $\log v < 4.00$, and using $H_0 = 55$ for larger
redshifts. The resulting distance moduli are in column (8). The
corresponding absolute magnitudes are in columns (9) to (11). The
decay rates $\Delta m_{15}(B)$ in column (12) are from the sources
listed by \cite{par99} (1999).
The data for the eight calibrators in Table~\ref{tbl5}, except for the
observed apparent magnitudes, are not shown in Table~\ref{tbl6}
because their absolute distance moduli are on the Cepheid system, not
based on redshifts.
The means of columns (9) to (11) for the absolute magnitudes in $B$,
$V$, and $I$ are shown at the foot of each of the three sections of
Table~\ref{tbl6}. The systematic progression, becoming fainter for
the earlier galaxy types, is evident and is significant at the 2-sigma
level. It is this difference that causes the separation of the ridge
lines in the Hubble diagram of spirals and E and S0 galaxies seen in
Fig.~10 of \cite{saha97} (1997).
However, this is not a type dependence per se. Note in column 12 that
the mean decay rates, shown at the foot of each section, differ
significantly between the three sections of Table~\ref{tbl6}. The
decay rates are much longer for the E to S0 types, averaging
$\langle{\Delta m_{15}(B)}\rangle = 1.44 \pm 0.04$ for these early
Hubble types, compared with $\langle{\Delta m_{15}(B)}\rangle = 1.03
\pm 0.03$ for the spirals in the first section of Table~\ref{tbl6}.
Because the decay rate itself is correlated with absolute magnitude
(next section), the {\it apparent} dependence of $\langle{M({\rm
max})}\rangle$ on Hubble type is in fact due to the decay rate
correlation. This, of course, is a clue as to differences in the
progenitor mass of the pre-SNe~Ia as a function of Hubble type, and
goes to the heart of the physics of the phenomenon (eg. \cite{nad98}
1998). In any case, the apparent correlation of $\langle{M}\rangle$
with Hubble type disappears when the decay rate corrections of the
next section are applied.
Said differently, the {\it apparent} dependence of mean absolute
magnitude with Hubble type is due to the difference in mean decay rate
between early type galaxies (E to S0 types) and late type spirals (the
first section of Table~\ref{tbl6}), together with the dependence of
decay rate on absolute magnitude at maximum (Fig.~\ref{fig12} in the
next section).
\subsection{The Decay Rate Dependence}
The absolute magnitudes in columns (9) to (11) of Table~\ref{tbl6} are
plotted in Fig.~\ref{fig12}. It is clear that there is a relation of
decay rate with (kinematic) absolute magnitude (\cite{phi93} 1993),
but that the relation is strongly non-linear. There is a clear
plateau between abscissa values of decay rates that are greater and
smaller than 1.0 to 1.5. In this plateau region, the absolute
magnitude is nearly independent of the decay rate.
Figure~\ref{fig12} shows the correlations of columns (9), (10), and
(11) (of Table~6) with the decay rate in column (12). The correlation
with absolute magnitude for small and for large $\Delta m_{15}(B)$ is
obvious. Nevertheless, the correlation is clearly non-linear. There
is a central region ($1.0 < \Delta m_{15}(B) < 1.5$) where the
correlation is weak. This is the core of the ``Branch normal''
majority of SNe~Ia, comprising 95\% of the observed SNe~Ia (only 1 in
20 of the local SNe~Ia discoveries are ``Branch abnormal''; see
\cite{bra93} 1993). It is only when the few known abnormal SNe~Ia are
added to the ``Branch normal'', i.e. blue SNe~Ia, that they show a
wide variation in luminosity.
Consider now the apparent correlation of $\langle{M_{\rm max}}\rangle$
with Hubble type, shown in Fig.~10 of \cite{saha97} (1997,
Paper~VIII). It is of central physical interest how the Hubble type
of the parent galaxy can produce slow-decay rate SNe~Ia in E galaxies,
and faster-decay rate SNe~Ia in later-type galaxies. One supposes
that the mass of the pre-SNe~Ia stars must be a function of chemical
evolution in any given parent galaxy, causing the change in the
character of the SNe~Ia explosion with Hubble type (\cite{nad98}
1998). Nevertheless, this is an astrophysical problem, not an
astronomical problem of how to use the data to determine reliable
distances. It suffices here to note that the type dependence on
$\langle{M_{\rm max}}\rangle$ disappears when the correction, now
described for decay rate (Fig.~\ref{fig12}), is applied.
The correlation of decay rate $\Delta m_{15}(B)$ with absolute
magnitudes in $B$, $V$, and $I$ in Fig.~\ref{fig12} is striking,
confirming Pskovskii's (1977, 1984) initial suggestions. But again,
the correlations are much less steep than suggested by \cite{phi93}
(1993) and by Hamuy et al.\ (1996b, their Fig.~2), especially in the
region of the ``Branch normal'' SNe~Ia with $0.95 < \Delta m_{15}(B) <
1.3$ where there is no strong correlation.
\cite{ham96a} (1994a) distinguish between the slope of the decay-rate
correlation with $M({\rm max})$ obtained from using TF, PN and SBF
distances to nearby SNe~Ia (as originally done by \cite{phi93} 1993),
versus that from using relative distances from redshifts at
intermediate distances. The latter sample gives a shallower slope than
that of Phillips. By restricting to a Branch normal sample, the slope
is further reduced.
In Fig.~\ref{fig12} we have fitted cubic polynomials to the data
contained in Table~\ref{tbl6} for the correlations of absolute
magnitudes in $B$, $V$, and $I$ in columns (9) to (11) with $\Delta
m(B)_{15}$ in column (12). Denote the correction to $M_i$ for decay
rate by $y_i$, such that the ``corrected'' absolute magnitude, reduced
to $\Delta m_{15}(B) = 1.1$, is defined by
\begin{displaymath}
M_i^{15} = M_i(~{\rm{Table~\ref{tbl6}}}) - y_i~.
\end{displaymath}
The correlations in Fig.~\ref{fig12}, based on the decay-rate data and
absolute magnitude data in Table~\ref{tbl6}, give the following cubic
corrections for decay rate vs. absolute magnitude in $B$, $V$, and
$I$. The lines drawn in Fig.~\ref{fig12} are calculated from these
equations. We reduce all data to $\Delta m_{15}(B) = 1.1$, and
therefore define $x = \Delta m_{15}(B) - 1.1$.
\begin{eqnarray}
y_B = 0.693x - 1.440x^2 + 3.045x^3\,, \\ y_V = 0.596x - 2.457x^2 +
4.493x^3\,, \\ y_I = 0.360x - 2.246x^2 + 4.764x^3\,.
\end{eqnarray}
Note that these polynomials are not well constrained outside the range
of $\Delta m(B)_{15}$ spanned by the SNe~Ia in this sample. In
particular, the upturn for very slow declining SNe~Ia is
uncertain. However, the cubic characterization given here is adequate
for the present arguments and calculations: it is not used in a region
where it is ill constrained.
The corrections for decay rate, $y_i$, from equations (15) to (17),
have been applied to columns (9) to (11) of Table~\ref{tbl6} to obtain
absolute magnitudes freed from the $\Delta m(B)_{15}$ parameter
effect. The arithmetic is not shown but the mean corrected magnitudes
$\langle{M_i^{15}}\rangle$, still binned into the three morphological
groups of Table~\ref{tbl6} are listed in the first four lines of
Table~\ref{tbl7}. The corresponding mean magnitudes
$\langle{M_B^{15}}\rangle$ and $\langle{M_V^{15}}\rangle$ of the eight
calibrators from Table~\ref{tbl5} are shown in the last line. No mean
value$\langle{M_I^{15}}\rangle$row is shown in the last line because
too few calibrators have known $I_{\rm max}$.
Table~\ref{tbl7} permits two conclusions:
(1) The type dependence shown in Fig.~10 of Paper~VIII between E
galaxies and spirals has now disappeared to within one sigma
differences after the $\Delta m_{15}$ corrections are applied.
(2) Comparing the $\langle{M_B^{15}}\rangle$ and
$\langle{M_V^{15}}\rangle$ values in the fourth line (the total
sample) with the corrected values of the calibrators in the fifth line
shows the statistical difference of $0\fm14 \pm 0\fm058$ in $B$ and
$0\fm15 \pm 0\fm076$ in $V$ between the fiducial sample and the
Table~\ref{tbl5} calibrators. Because the absolute magnitudes of the
fiducial sample are based on $H_0 = 55$, these data, corrected for
decay rate, give
\begin{equation}
H_0(B)(~{\rm decay~rate}) = 58.8 \pm 2\ksm\,,
\end{equation}
and
\begin{equation}
H_0(V)(~{\rm decay~rate}) = 59.1 \pm 2\ksm\,.
\end{equation}
It is shown in \cite{par99} (1999), that the effect of the decline
rate correction taken alone, is to increase $H_0$ by 7\%. This
correction depends on the extent to which the decline rate
distribution is different for the calibrating and distant samples of
SNe~Ia.
\subsection{The Color Dependence}
Figure~\ref{fig13} shows that there is still a dependence of the
corrected $M_i^{15}$ magnitudes on color that is not removed by the
$\Delta m_{15}$ corrections, contrary to the removal of the dependence
on Hubble type. The symbols are the same as in Fig.~\ref{fig12}.
The ordinates are the kinematic ($H_0 = 55$) absolute magnitudes,
$M_i^{15}$, based on variable local values of $H_0$ and the decay rate
corrections from the last subsection. The abscissa is the observed
$(B_0 - V_0)$ color that can be derived from columns (4) and (5) of
Table~\ref{tbl6}.
The equations for the least squares solutions for the lines are
\begin{equation}
M_B^{15} = 1.712\cdot (B_0 - V_0) - 19.648\,, \\
\end{equation}
\begin{equation}
M_V^{15} = 0.587\cdot (B_0 - V_0) - 19.562\,,
\end{equation}
and
\begin{equation}
M_I^{15} = -0.216\cdot (B_0 - V_0) - 19.312\,.
\end{equation}
Figure~\ref{fig14} is the same as Fig.~\ref{fig13} but with $(V_0-
I_0)$ colors derived from columns (10) and (11) of
Table~\ref{tbl6}. The least squares solutions for the lines are
\begin{equation}
M_B^{15} = 0.257\cdot (V_0 - I_0) - 19.540\,, \\
\end{equation}
\begin{equation}
M_V^{15} = -0.260\cdot (V_0 - I_0) - 19.621\,,
\end{equation}
and
\begin{equation}
M_I^{15} = -1.115\cdot (V_0 - I_0) - 19.607\,.
\end{equation}
The immediate consequence of equations (20) to (25) is that these
color dependencies {\it are not due to reddening and absorption}
because the slope coefficients in $B$ and $V$, nor their ratios,
conform to the canonical normal interstellar values of $A_B/E(B-V) =
4$, $A_V/E(B-V) = 3$, and $A_I/E(B-V) = 1.76$, and $A_B/E(V-I)= 3.2$,
$A_V/E(V-I) = 2.4$, and $A_I/E(V-I) = 1.4$ (\cite{schef82} 1982). If,
for example, the variation of absolute magnitude with color index from
SN to SN were due entirely to absorption and reddening, then the slope
coefficients in the correlations of Fig.~\ref{fig13}, (Eqs.\,20 to 22)
would have to be $dM_B/E(B-V) = 4$, $dM_V/E(B-V) = 3$, and
$dM_I/E(B-V) = 1.76$, based on $A_V/A_I = 1.7$. Not only are the
coefficients for $B$ and $V$ in equations (20) and (21) very different
from these requirements for reddening, but even the sign of the
variation in $I$ with $(B-V)$ is negative. This is impossible if the
cause is internal reddening with its corresponding dimming.
The situation is even more decisive from equations (23) to (25) using
$(V-I)$ colors in Fig.~\ref{fig14}. From E(V-I)/E(B-V) = 1.25 (Eq.\,1
of \cite{dea78} 1978), the predictions are $dM_B/E(V-I) = 3.2$,
$dM_V/E(V-I) = 2.4$, and $dM_I/E(V-I) = 1.41$. These are not only far
different from the coefficients in equations (23) to (25), but again
the sense of the observed correlation in equations (24) and (25)
belies a reddening/extinction explanation. {\it Brighter} magnitudes
for redder colors are not possible for any known reddening and
absorption law.
The conclusion is that the correlations in Figs.~\ref{fig13} and
\ref{fig14} are due to intrinsic properties internal to the physics of
the SNe themselves rather than to reddening and absorption. Hence,
any corrections to the absolute magnitudes based solely on assumed
absorption-to-reddening ratios for normal extinction are most likely
to be incorrect.
Nevertheless, the correlations in Figs.~\ref{fig13} and \ref{fig14}
are definite, and can be used to further reduce the absolute
magnitudes to some fiducial value of observed color. We thus reduce
to a fiducial color of $(B_0-V_0) = 0.00$ as a second parameter, in
addition to the decay-rate parameter, in agreement with \cite{tri98}
(1998).
Our approach is different in principle from that of \cite{phi99}
(1999), where they assert that the color range at given decline rate
is due to reddening, and proceed accordingly. Our conclusion from the
above analysis is that for our sample of SNe~Ia, with $(B_{\rm max} -
V_{\rm max}) < 0.20$, the color spread at given decline rate is
intrinsic in nature, since correlation of the peak brightness with the
residual color for this sample is different from what is expected from
reddening.
Applying the color-term corrections of equations (20) and (21) to the
$M_B$ and $M_V$ magnitudes of Table~\ref{tbl6}, and taking the mean
values over the complete sample of 34 SNe~Ia in Table~\ref{tbl6} where
the decay rates are known, gives
\begin{equation}
\langle{M_B^{\rm corr}}\rangle = \langle{M_B^{15} + ~{\rm
color~term}}\rangle = -19.64 \pm 0.026\,,
\end{equation}
and
\begin{equation}
\langle{M_V^{\rm corr}}\rangle = \langle{M_V^{15} + ~{\rm
color~term}}\rangle = -19.61 \pm 0.026\,.
\end{equation}
These magnitudes are calculated with variable $H_0$ following the
precepts of \S 6.1 and reduced to $\Delta m_{15} = 1.1$ and $(B_0 -
V_0)= 0.00$.
For the calibrators in Table~\ref{tbl5} one finds in analogy
\begin{equation}
\langle{M_B^{\rm corr}}\rangle = -19.45 \pm 0.060\,,
\end{equation}
and
\begin{equation}
\langle{M_V^{\rm corr}}\rangle = -19.44 \pm 0.079\,.
\end{equation}
Comparison of equations (28) and (29) with equations (26) and (27)
shows that the $H_0 = 55$ assumption must be changed to accommodate
the differences of $0\fm19 \pm 0\fm08$ in $B$ and $0\fm17 \pm 0\fm09$
in $V$. They require the global value of $H_0$, now fully corrected
for the three effects of (1) the change of $H_0$ outward, (2) the
light curve decay rate, and (3) the color variation with $M({\rm
max})$, to be
\begin{equation}
H_{0}(B) = 60.2 \pm 2\ksm\,,
\end{equation}
and
\begin{equation}
H_{0}(V) = 59.6 \pm 2\ksm\,.
\end{equation}
These values are only 2\% higher than equations (18) and (19) which
use only the decay-rate correction, and 1\% to 4\% higher than
equations (12) and (13) which are weighted towards SNe~Ia in spirals.
Other authors have adopted $\langle{M({\rm max})}\rangle$ values that
differ from those in equations (10) and (11) for the calibrator
absolute magnitudes. For example, \cite{ken98} (1998), and
\cite{fre99} (1999) have discarded several of our calibrators, and
added two that are not based on direct Cepheid distances to the host
galaxy. They have consequently derived a fainter mean absolute $B$
magnitude than in equation (10). Specifically, they assume that the
distances of the early-type galaxies NGC~1316 and NGC~1380 in the
Fornax cluster, parent galaxies to SN~1980N and SN~1992A, are
identical with that of the spiral NGC~1365 for which there is a
Cepheid distance. \cite{sun98} (1998) have also considered the
questionable SN~1980 and SN~1992A as possible calibrators.
However, there are reasons to suspect that NGC~1365 is in the
foreground of the Fornax cluster (\cite{satasa99} 1999), and therefore
that the precept of the fainter calibration used by \cite{ken98}
(1998), calibrating the two Fornax SNe~Ia via NGC~1365, is not
correct. The evidence is that \cite{wel94} (1994) have demonstrated
that the multi-color light curves of SN~1989B in NGC~3627 and SN~1980N
in NGC~1316 are virtually identical, and in fact establish the
reddening and extinction to SN~1989B by comparing the magnitude shifts
in different passbands relative to SN~1980N. Asserting then that
SN~1980N has the same peak brightness as SN~1989B yields the distance
modulus difference of $1\fm62 \pm 0\fm03$ (\cite{wel94} 1994). There
is additional uncertainty of $\pm 0.17$ to allow for scatter in the
difference in peak brightness of two SNe~Ia with the same decline
rate. With our modulus of $(m - M)_0 = 30.22 \pm 0.12 $ (Eq.~7) for
SN~1989B, the derived modulus of NGC~1316, host to SN~1980N, is $31.84
\pm 0.21$. This is $0\fm5$ more distant than the Cepheid distance of
NGC~1365 (\cite{mad98} 1998), but is close to the value found for the
early-type galaxies of the Fornax cluster by independent methods
(\cite{tam98b} 1998b). In any case the implication of \cite{ken98}
(1998) and \cite{fre99} (1999) that the two SNe~Ia 1989B and 1980N
differ by $0\fm5$ in luminosity, which is based on the unproved
assertion that NGC~1365 and NGC~1316 are at the same distance, is not
credible.
The further consequence of the precept of equating the Cepheid
distance of NGC~1365 to that of the unknown distance to NGC~1316 and
NGC~1380 is that the decay-rate absolute magnitude relation which
\cite{fre99} (1999) deduce using their faint absolute magnitudes of
the two Fornax SNe~Ia is steeper than that in Fig.~\ref{fig12} by an
amount that compromises their conclusions concerning $H_0$, explaining
their abnormally high value of $H_0 \approx 73$.
\section{Summary and Conclusions}
(1) Comparison of 12 epoch $HST$ frames in the $F555W$ band and
five epoch frames in the $F814W$ band has isolated 68 definite Cepheid
variables with periods between 3 and $\sim75$ days in the dust-rich
galaxy NGC~3627 in the Leo Group.
(2) The adopted true modulus of $(m - M)_0 = 30.22 \pm 0.12$ that
results from analysis of various subsamples of the data according to
the degree of internal absorption, agrees well with the distance
moduli of $(m - M)_0 = 30.37 \pm 0.16$ for NGC~3368 (parent of
SN~1998bu) (\cite{tan95} 1995) and $(m - M)_0 = 30.01 \pm 0.19$ for
NGC~3351 (\cite{gra97} 1997), showing that the extended Leo Group in
fact exists (\cite{hum56} 1956, Table XI).
(3) Combining the NGC~3627 Cepheid modulus of $(m -M)_0 = 30.22$
with the photometry of \cite{wel94} (1994) for the daughter SN~Ia
SN~1989B (corrected for extinction) gives the absolute magnitudes in
$B$ and $V$ at maximum light of $M_B = -19.36 \pm 0.18$, and $M_V =
-19.34 \pm 0.16$ for SN~1989B as listed in Table~\ref{tbl5}.
(4) Combining these absolute magnitudes with those for the
previous six calibrators determined in previous papers of this series,
and with two additional calibrators recently available (SN~1974G in
NGC~4414 and SN~1998bu in NGC~3369), gives a mean calibration (without
second parameter corrections) of $\langle{M_B}\rangle = - 19.49 \pm
0.07$ and $\langle{M_V}\rangle = - 19.48 \pm 0.07$.
(5) The absolute magnitudes $M_{\rm max}$ of blue SNe~Ia with
$(B_0 - V_0) <0.20$ correlate with the light curve decay rate $\Delta
m_{15}$. The correlation is approximated by a cubic equation which has
a flat plateau at intermediate values of $\Delta m_{15}$. In
addition, $M_{\rm max}$ depends on the color $(B_0 - V_0)$. For the
derivation of the dependence on these second parameters, a slight
decrease of $H_0$ out to $10,000 \kms$ has been taken into
account. Once corrections for these two second parameters are applied,
no residual dependence of $M_{\rm max}$ on Hubble type is seen.
(6) The calibrators have somewhat smaller mean values of $\Delta
m_{15}$ and bluer mean colors $(B_0 - V_0)$ than the distant SNe~Ia
defining the Hubble diagram. The application of second parameter
corrections therefore tends to increase $H_0$, but the effect is less
than 10\%.
(7) The overall value of the Hubble constant determined here from
the current eight calibrators as applied to the distant SNe~Ia with
$1000 \kms < v < 30,000 \kms$ is $H_0 = 60 \pm 2$ in both $B$ and $V$
(Eqs. 30 and 31).
(8) The Hubble constant derived here rests on Cepheid distances
whose zeropoint is set to an LMC modulus of 18.50. If the latter is
revised upwards by $\approx 0\fm06$ (\cite{fed98} 1998), the
consequent value of the Hubble constant is
\begin{equation}
H_0 = 58 \pm 2\ksm\,.
\end{equation}
\acknowledgments{Acknowledgment}
We thank the many individuals at STScI who worked hard behind the
scenes to make these observations possible, and wish to particularly
mention Doug van Orsow, George Chapman, Bill Workman, Merle Reinhardt
and Wayne Kinzel. A.S. and A.S. acknowledge support from NASA through
grant GO-5427.02-93A from the Space Telescope Science Institute, which
is operated by the Association of Universities for Research in
Astronomy. L.L. and G.A.T. thank the Swiss National Science
Foundation for continued support.
\clearpage
|
\section{Introduction}
Two photon reactions are an important part of physics which is being
studied in current $e^+e^-$ experiments at LEP1 and LEP2 and which will also
be intensively analyzed in future $e^+e^-$ colliders.
The available photon-photon energy and photon virtualities continously
increase with the increasing energy of the $e^+e^-$ pair. Therefore the
data from LEP1 and LEP2 and the expected results from the TESLA and NLC
provide us with an excellent oportunity to study virtual photon
scattering in the diffractive regime. Moreover, with proper
experimental cuts, it is possible to study observables dominated by the
perturbative QCD contributions. The theoretical description of such
processes is based on expectations concerning high energy limit in perturbative QCD
which is at present theoretically fairly well understood \cite{GLR,LIPAT1}. The
leading high energy behaviour is controlled by the pomeron singularity which
corresponds to the sum of ladder diagrams with reggeized gluons along
the chain. This sum is described by the Balitzkij, Fadin, Kuraev,
Lipatov (BFKL) equation \cite{BFKL1}.
The perturbative QCD pomeron exchange effects can be observed only in
specific conditions and even then not in the unambigous form.
In order to minimize the contribution of the other mechanisms competing with
the QCD pomeron and to guarantee the validity of the calculations based on
perturbative QCD one has to chose carefully the processes to analyze.
The virtualities of the gluons along the
ladder should be large enough to assure the applicability of the
perturbative expansion. The neccesary hard scale may be provided either
by coupling of the ladder to scattering particles, that contain a hard
scale themselves, or by large momentum transfer carried by the gluons.
Moreover, to distinguish the genuine BFKL from DGLAP evolution effects
it is convenient to focus on procesess in which the scales on both
ends of the ladder are of comparable size. Finally, one requires that the
non-perturbative effects should factor out in order to minimize the theoretical
uncertainties.
The two classical processes which can probe the
QCD pomeron in $ep$ and in $\gamma^* p$ collisions are the deep inelastic events
accompanied by an energetic (forward) jet {\cite{MUELLERJ, DISJET}
and the production of large $p_T$ jets separated by the rapidity gap
\cite{JETGAP}. The former process probes the QCD pomeron in the
forward direction while the latter reflects the elastic scattering of
partons via the QCD pomeron exchange with non-zero (and large)
momentum transfer. Another possible probe of the QCD pomeron at
(large) momentum transfers can be provided by the diffractive vector
meson photoproduction accompanied by proton dissociation in order to
avoid nucleon form-factor effects \cite{FORSHAW,BARTLQ}.
In this talk we shall analyze two measurements in $e^+ e^-$ collisions,
complementary to those listed above.
Namely we focus on double diffractive $J/\psi$ production in
$\gamma\gamma$ collisions and on the total $\gamma^* \gamma^*$ cross section.
The former process is unique since in principle it allows to test the
QCD pomeron for arbitrary momentum transfers \cite{KMPSI}.
The hard scale is given by the relatively large mass of the $c$-quark.
The total $\gamma^*\gamma^*$ cross-section has been studied by several
authors \cite{GGSTAR,BRODSKY}, however
our approach has the novel feature of taking into account dominant non-leading
corrections to the BFKL equation. This re-analysis has become necessary when
the next-to-leading corrections to the BFKL kernel were obtained
\cite{BFKLNL}, which alter substantially the results obtained at the leading
order.
It turns out that the magnitude of the next-to-leading (NLO),
i.e. $O(\alpha_s^2)$,
contribution to the QCD pomeron intercept is very large for the values of the
QCD coupling within the range which is relevant for most experiments. This
means that the NLO approximation alone is not reliable and one has to perform
resummation to all orders. Unfortunately the exact result of this resummation is
unknown.
It may however be possible to pin down certain dominant contributions of well
defined physical origin and perform their exact resummation \cite{RESUM,KMSG}.
In our approach we shall use the so called consistency constraint which limits
the available phase space for the real gluon emission by imposing the requirement
that the virtuality of the exchanged gluons along the chain is dominated by their
transverse momentum squared. Let us remind that the form of the LO BFKL kernel
where the gluon propagators contain only the gluon transverse momentum squared
etc. is only valid within the region of phase space restricted by this
constraint. Formally however, the consistency constraint generates
subleading corrections. It can be shown that at the NLO accuracy it generates
about 70 \% of the exact result for the QCD pomeron intercept. The very important
merit of this constraint is also the fact that it automatically generates
resummation of higher order contributions which stabilizes the solution
\cite{KMSG}.
\section{The total ${\gamma^* \gamma^*}$ cross-section}
The collisions of virtual photons may be studied experimentally only
as subprocesses of reactions between charged particles. In
principle, one is able to unfold the photonic cross-section from the
leptonic data, however this procedure requires additional assumptions which
increase the systematic uncertainty of the result. It seems to be
more sensible to formulate the predictions for the $e^+e^-$ cross-sections
with the properly chosen cuts and compare them directly with the $e^+ e^-$ data.
Therefore we use the equivalent photon approximation which allows us to
express the leptonic cross-section through a convolution of the photonic
cross-section and the standard flux factors.
Thus the cross-section for the process $e^+e^- \rightarrow e^+e^- + X$
(averaged over the angle $\phi$ between the lepton scattering planes in
the frame in which the virtual photons are aligned along the $z$ axis)
is given by the following formula \cite{BRODSKY}:
$$
{Q_1^2 Q_2^2 d\sigma \over dy_1 dy_2 dQ_1^2 dQ_2^2} =
\left({\alpha\over 2 \pi}\right)^2
[P^{(T)}_{\gamma/e^+}(y_1)P^{(T)}_{\gamma/e^-}(y_2)
\sigma^{TT}_{\gamma^* \gamma^*}(Q_1^2,
Q_2^2,W^2)+
$$
$$
P^{(T)}_{\gamma/e^+}(y_1)P^{(L)}_{\gamma/e^-}(y_2)
\sigma^{TL}_{\gamma^* \gamma^*}(Q_1^2,Q_2^2,W^2)+
P^{(L)}_{\gamma/e^+}(y_1)P^{(T)}_{\gamma/e^-}(y_2)
\sigma^{LT}_{\gamma^* \gamma^*}(Q_1^2,Q_2^2,W^2)+
$$
\begin{equation}
P^{(L)}_{\gamma/e^+}(y_1)P^{(L)}_{\gamma/e^-}(y_2)
\sigma^{LL}_{\gamma^* \gamma^*}(Q_1^2, Q_2^2,W^2)]
\label{conv}
\end{equation}
where
\begin{equation}
P^{(T)}_{\gamma/e}(y) = {1 + (1-y)^2\over y}
\label{pt}
\end{equation}
\begin{equation}
P^{(L)}_{\gamma/e}(y) = 2{1-y\over y}
\label{pl}
\end{equation}
where $y_1$ and $y_2$ are the longitudinal momentum fractions of the parent
leptons carried by virtual photons, $Q_i^2 = -q_i^2$
($i=1,2$) where $q_{1,2}$ denote the four momenta of the
virtual photons and $W^2$ is the total CM energy squared of the
two (virtual) photon system, i.e. $W^2=(q_1+q_2)^2$. The cross-sections
$\sigma^{ij}_{\gamma^* \gamma^*}(Q_1^2,Q_2^2,W^2)$ are the total
cross-sections for the process $\gamma^* \gamma^* \rightarrow X$
and the indices $i,j=T,L$ denote the polarization of
the virtual photons. The functions $P^{(T)}_{\gamma/e}(y)$ and
$P^{(L)}_{\gamma/e}(y)$ are the transverse and longitudinal photon
flux factors.
\noindent
\begin{figure}
\leavevmode
\begin{center}
\parbox{6.5cm}{
{\large a)}\\
\epsfxsize = 5.5cm
\epsfysize = 4.2cm
\epsfbox[60 545 311 741]{ggfig1.eps}}\qquad
\parbox{6.5cm}{
{\large b)}\\
\epsfxsize = 5.5cm
\epsfysize = 4.2cm
\epsfbox{ppfig1.eps}} \\
\end{center}
\caption{\small The QCD pomeron exchange mechanism of the processes
a) $\gamma_1 ^* (Q_1 ^2) \gamma_2 ^* (Q_2 ^2) \to X$ and b)
$\gamma\gamma \rightarrow J/\psi J/\psi$.}
\end{figure}
The ladder diagram corresponding to the perturbative contribution to
the diffractive subprocess $\gamma_1 ^* (Q_1 ^2) \gamma^* (Q_2 ^2) \to X$
is shown in Fig.~1a.
The cross-sections $\sigma^{ij}_{\gamma^* \gamma^*}(Q_1^2,Q_2^2,W^2)$
are given by the following formulae:
$$
\sigma^{ij}_{\gamma^* \gamma^*}(Q_1^2,
Q_2^2,W^2) = P_S(Q_1^2,Q_2^2,W^2)\delta_{iT}\delta_{jT} +
$$
\begin{equation}
{1\over 2 \pi}\sum_q\int_{k_0^2}^{k_{max}^2(Q_2^2,x)} {d^2k\over \pi k^4}
\int _{\xi_{min}(k^2,Q_2^2)}^{1/x} d\xi
G^{0j}_q(k^2,Q_2^2,\xi)
\Phi_i(k^2,Q_1^2,x\xi)
\label{csx}
\end{equation}
where
\begin{equation}
k_{max}^2(Q_2^2,x)=-4m_q^2+Q_2^2 \left( {1\over x} - 1 \right)
\label{kmax}
\end{equation}
\begin{equation}
\xi_{min}(k^2,Q^2)=1+{k^2+4m_q^2\over Q^2}
\label{ximin}
\end{equation}
and
\begin{equation}
x={Q_2^2\over 2q_1q_2}
\label{x}
\end{equation}
In Eq.~(\ref{csx}) we sum over four quark flavours with $m_q \to 0$ for
light quarks and $m_c=1.5\rm \; GeV$.
The lower limit of integration over $k^2$ appearing in Eq.~(\ref{csx})
is taken to be $k_0 ^2 = 1\rm \; GeV ^2$ in order to subtract the contribution
from the nonperturbative region from the perturbative part of the
amplitude.
The functions $G^{0i}_q(k^2,Q^2,\xi)$ are defined as
below: \cite{BRODSKY,KMSTAS}
$$
G^{0T}_q(k^2,Q^2,\xi)=
$$
$$
2 \alpha_{em} \alpha_s(k^2+m_q^2)e_q^2\int_0^{\lambda_{max}} d \lambda
\int {d^2p^{\prime }\over \pi}\;
\delta\left[\xi-\left(1+{p^{\prime 2}+m_q^2\over z(1-z)Q^2} +
{k^2\over Q^2}\right)\right] \times
$$
\begin{equation}
\left\{ \left[(z^2 + (1-z)^2)\left({\tdm p\over D_1} -
{\tdm p + \tdm k\over D_2} \right)^2 \right]
+m_q^2 \left( {1\over D_1} - {1\over D_2} \right)^2 \right\}
\label{g0t}
\end{equation}
$$
G^{0L}_q(k^2,Q^2,\xi)=
$$
$$
8 \alpha_{em} \alpha_s(k^2+m_q ^2)e_q^2\int_0^{\lambda_{max}} d \lambda
\int {d^2p^{\prime }\over \pi}\;
\delta\left[\xi-\left(1+{p^{\prime 2}+m_q^2\over z(1-z)Q^2} +
{k^2\over Q^2}\right)\right] \times
$$
\begin{equation}
\left[z^2 (1-z)^2 \, \left({1\over D_1} -
{1\over D_2} \right)^2 \right]
\label{g0l}
\end{equation}
where
\begin{equation}
z={1+\lambda\over 2}
\label{zlam}
\end{equation}
\begin{equation}
\tdm p=\tdm p^{\prime} + (z-1) \tdm k
\label{pprime}
\end{equation}
$$
D_1=p^2 + z(1-z)Q^2 + m_q^2
$$
\begin{equation}
D_2=(p+k)^2 + z(1-z)Q^2 + m_q^2
\label{d12}
\end{equation}
In the formulae given above as well as throughout the rest of the
text we are using the one loop approximation for the QCD coupling
$\alpha_s$ with the number of flavours $N_f=4$ and set
$\Lambda_{QCD}=0.23 {\rm \; GeV}$.
The function $P_S(Q_1^2,Q_2^2,W^2)$ corresponds to the
contribution from the region $k^2 \le k_0^2$ in the
corresponding integrals over the gluon transverse momenta. It
is assumed to be dominated by the soft pomeron contribution
which is estimated from the factorisation of its couplings, i.e.
\begin{equation}
P_S(Q_1^2,Q_2^2,W^2) = {\sigma^{SP}_{\gamma^*(Q_1^2)p}(Q_1^2,W^2)
\sigma^{SP}_{\gamma^*(Q_2^2)p}(Q_2^2,W^2)\over \sigma_{pp}^{SP}}
\label{sp}
\end{equation}
We assume that this term is only contributing to the transverse
part. In equation (\ref{sp}) the cross-sections
$\sigma^{SP}_{\gamma^*(Q_i^2)p}(Q_i^2,W^2)$ and
$\sigma_{pp}^{SP}$ are the soft pomeron contributions to
the $\gamma^*p$ and $pp$ total cross sections and their
parametrisation is taken from Refs. \cite{DLTOTCX,DLDIS}. Their
$W^2$ dependence is, of course, universal i.e.
$$
\sigma_{pp}^{SP}=\beta_p^2\left({W^2\over W_0^2}\right)^{\alpha_{SP}(0)-1}
$$
\begin{equation}
\sigma^{SP}_{\gamma^*(Q_i^2)p}(Q_i^2,W^2) = \beta_{\gamma^*}(Q^2)\beta_p
\left({W^2\over W_0^2}\right)^{\alpha_{SP}(0)-1}
\label{sppar}
\end{equation}
with $W_0 = 1\rm \; GeV$ and $\alpha_{SP}(0) \approx 1.08$.
The function $ \Phi_T(k^2,Q^2,x_g) $ satisfies the Balitzkij,
Fadin, Kuraev, Lipatov (BFKL) equation which, in the leading
$\ln (1/x)$ approximation has the following
form:
$$
\Phi_i(k^2,Q^2,x_g)=\Phi^0_i(k^2,Q^2,x_g)+\Phi^S (k^2,Q^2,x_g)\delta_{iT}+
{3\alpha_s(k^2)\over \pi} k^2\int_{x_g}^1 {dx^{\prime}\over x^{\prime}}
\int_{k_0^2}^{\infty} {dk^{\prime 2} \over k^{\prime 2}}
$$
\begin{equation}
\left [ {\Phi_i(k^{\prime 2},Q^2,x^{\prime}) - \Phi_i(k^{ 2},Q^2,x^{\prime})
\over |k^{\prime 2} - k^{ 2}|} + {\Phi_i(k^{ 2},Q^2,x^{\prime})\over
\sqrt{4 k^{\prime 4} + k^{4}}}\right]
\label{bfklll}
\end{equation}
In what follows we shall consider the modified BFKL equation in
which we restrict the available phase-space in the real gluon
emission by the consistency constraint:
\begin{equation}
k^{\prime 2} \le k^2{ x^{\prime}\over x_g}
\label{cc}
\end{equation}
This constraint follows from the requirement that the virtuality
of the exchanged gluons is dominated by their transverse momentum
squared. The consistency constraint (\ref{cc}) introduces the
non-leading $\ln(1/x)$ effects and in the next-to-leading
approximation exhausts about 70\% of the entire next-to-leading corrections
to the QCD pomeron intercept. The modiffied BFKL equation takes the
following form:
$$
\Phi_i(k^2,Q^2,x_g)=\Phi^0_i(k^2,Q^2,x_g)+\Phi^S(k^2,Q^2,x_g)\delta_{iT}+
{3\alpha_s(k^2)\over \pi} k^2\int_{x_g}^1 {dx^{\prime}\over x^{\prime}}
\int_{k_0^2}^{\infty} {dk^{\prime 2} \over k^{\prime 2}}
$$
\begin{equation}
\left [ {\Phi_i(k^{\prime 2},Q^2,x^{\prime})\Theta
\left(k^2{ x^{\prime}\over x_g}
-k^{\prime 2}\right) - \Phi_i(k^{ 2},Q^2,x^{\prime})
\over |k^{\prime 2} - k^{ 2}|} + {\Phi_i(k^{ 2},Q^2,x^{\prime})\over
\sqrt{4 k^{\prime 4} + k^{4}}}\right]
\label{bfklcc}
\end{equation}
The inhomogeneous terms in equations (\ref{bfklll}, \ref{bfklcc})
are the sum of two contributions $\Phi^0_i(k^2,Q^2,x_g)$
and $\Phi^S(k^2,Q^2,x_g)\delta_{iT}$. The first term
$\Phi^0 _i (k^2, Q^2, x_g)$
corresponds to the diagram in which the two gluon system couples
to a virtual photon through a quark box and are given by
following equations:
\begin{equation}
\Phi^0_i(k^2,Q^2,x_g)=\sum_q \int_{x_g}^1 dz \; \tilde G^0_{iq}(k^2,Q^2,z)
\label{phii0}
\end{equation}
where
$$
\tilde G^0_{Tq}(k^2,Q^2,z)=2\alpha_{em} e_q^2\alpha_s(k^2+m_q^2)
\int_0^1 d\lambda \left\{
{[\lambda^2 + (1-\lambda)^2][z^2+(1-z)^2] k^2 \over \lambda(1-\lambda)k^2+
z(1-z)Q^2 + m_q^2} \right. +
$$
\begin{equation}
2m_q^2 \left. \left[ {1\over z(1-z)Q^2 + m_q^2} -
{1\over \lambda(1-\lambda)k^2+z(1-z)Q^2 + m_q^2}\right] \right\}
\label{tgt0}
\end{equation}
$$
\tilde G^0_{Lq}(k^2,Q^2,z)=16\alpha_{em}Q^2 k^2 e_q^2\alpha_s(k^2+m_q^2)
\times
$$
\begin{equation}
\int_0^1d\lambda
\left\{ {[\lambda (1-\lambda)][z^2(1-z)^2] \over [\lambda(1-\lambda)k^2+
z(1-z)Q^2 + m_q^2][z(1-z)Q^2 + m_q^2]} \right\}
\label{tgl0}
\end{equation}
The second term $\Phi^S(k^2,Q^2,x_g)\delta_{iT}$, which is
assumed to contribute only to the transverse component,
corresponds to the contribution to the BFKL equation from the
nonperturbative soft region $k^{\prime 2} < k_0^2$. Adopting
the strong ordering approximation $k^{\prime 2} \ll k^2$ it
is given by the following formula:
\begin{equation}
\Phi^S (k^2,Q^2,x_g)=
{3 \alpha_s(k^2)\over \pi}
\int_{x_g}^1 {dx^{\prime}\over x^{\prime}}
\int_{0}^{k_0^2} {dk^{\prime 2} \over k^{\prime 2}}
\Phi_T(k^{\prime 2},Q^2,x^{\prime})
\label{stophi}
\end{equation}
The last integral in equation (\ref{stophi}) can be interpreted
as a gluon distribution in a virtual photon of virtuality $Q^2$
evaluated at the scale $k_0^2$. At low values of $x^{\prime}$
it is assumed to be dominated by a soft pomeron contribution and
can be estimated using the factorisation of the soft pomeron
couplings:
\begin{equation}
\int_{0}^{k_0^2} {dk^{\prime 2} \over k^{\prime 2}}
\Phi_T(k^{\prime 2},Q^2,x^{\prime})=\pi^2 x^{\prime}g_p(x^{\prime},k_0^2)
{\beta_{\gamma^*}(Q^2)\over \beta_p}
\label{inhos}
\end{equation}
where $g_p(x^{\prime},k_0^2)$ is the gluon distribution in a
proton at the scale $k_0^2$
and the couplings $\beta_{\gamma^*}(Q^2)$ and $\beta_p$
are defined by equation (\ref{sppar}).
We adopt the parametrization of the gluon
structure function taken from Ref.\cite{KMSTAS} i.e.
$xg(x,k_0 ^2) = 1.57 (1-x)^{2.5}$ which is consinstent with the DIS data.
\noindent
\begin{figure}[hbpt]
\begin{center}
\epsfxsize = 13cm
\epsfysize = 13cm
\epsfbox[18 260 555 775]{ggtot.ps}
\end{center}
\caption{\small
Energy dependence of the cross-section
$\sigma^{TT} _{\gamma^* \gamma^*}(Q_1 ^2, Q_2 ^2, W^2)$
for the process $\gamma^*(Q_1 ^2) \gamma^*(Q_2^2) \rightarrow X$
for various choices of virtualities
$Q^2 = Q_1 ^2 = Q_2 ^2$ corresponding to Eq.~(\ref{csx}).
For each choice of the virtuality
four curves are shown taking into account hard effects only (``hard part''),
hard amplitude with soft pomeron contributions added in the source term
of the BFKL equation (``mixed''),
the full cross-section including both soft and hard pomeron
contributions (``full result''). We also show the ``full result'' with the
low scale of $\alpha_s$ in the impact factors:
$\mu^2 = (k^2 + m_q^2)/4$.
}
\end{figure}
In Fig.~2 we show our results for
$\sigma^{TT}_{\gamma^* \gamma^*} (Q_1 ^2, Q_2 ^2, W^2)$ plotted
as the function of the CM energy $W$ for three different values of $Q^2$
where $Q_1^2 = Q_2^2 = Q^2$. We plot in this figure:
\begin{enumerate}
\item the pure QCD
(i.e. ``hard'') contribution obtained from solving the BFKL equation with the
consistency constraint included (see Eq.~(\ref{bfklcc})) and with the
inhomogeneous term containing only the QCD impact factor defined by equations
(\ref{phii0},\ref{tgt0},\ref{tgl0}),
\item
the ``mixed" contribution generated by the BFKL equation (\ref{bfklcc})
with the soft pomeron contribution defined by equations (\ref{stophi},
\ref{inhos}) included in the inhomogeneous term,
\item
The ``full" contribution which also contains the soft pomeron term (\ref{sp}).
\end{enumerate}
We also show results obtained by changing the scale of the strong coupling
$\alpha_s$ in the impact factors from $k^2+m_q^2$ to $(k^2+m_q^2)/4$.
The scale of $\alpha_s$ in the BFKL equation is the same in the both cases.
The components of the cross-section for which at least one of the photons
is longitudinally polarized have very similar energy dependence to
$\sigma^{TT} _{\gamma^* \gamma^*} (Q_1^2,Q^2_2,W^2)$
and give together about 60\% of the transverse-transverse contribution.
We see from this figure that the effects of the soft pomeron contribution
are non-negligible at low and moderately large values of $Q^2 < 10 \rm \; GeV^2$
and for moderately large values of $W < 100 \rm \; GeV$.
The QCD pomeron however dominates already at $Q^2=40 \rm \; GeV^2$.
We also see from this figure that for low energies $W<40 \rm \; GeV$ the
phase-space effects are very important.
For $W>40\rm \; GeV$ or so one observes that the cross-section exhibits
the effective power-law behaviour
$\sigma_{\gamma^* \gamma^*}(W) \sim (W^2)^{\lambda_P}$.
The (effective) exponent increases weakly
with increasing $Q^2$ and varies from $\lambda_P=0.28$ for $Q^2=2.5\rm \; GeV^2$ to
$\lambda_P=0.33$ for $Q^2 = 40 \rm \; GeV^2$. This (weak) dependence of the
effective exponent $\lambda_P$ with $Q^2$ is the result of the interplay
between soft and hard pomeron contributions,
where the former becomes less important at large $Q^2$.\\
\begin{table}
\caption{ Comparison of the theoretical results to L3 data for $e^+ e^- \to e^+ e^- X$
with $E_{tag} > 30$ GeV, 30 mrad $ < \theta_{tag} <$ 66 mrad. We show in
the table $d\sigma / dY$ binned in $Y$ obtained from experiment and
the results of our calculation which take into account perturbative pomeron only
(hard) and both perturbative and soft pomerons (hard + DL) for two different
choices of scale of the $\alpha_s$ in impact factors and for $e^+e^-$ CM
energy 91~GeV and 183~GeV.}
\begin{center}
\begin{tabular}{||c|c||c|c||c|c||}
\hline\hline
& \multicolumn{5}{|c||}{$\langle d\sigma / dY \rangle$ [fb] }\\
\cline{2-6}
& & \multicolumn{4}{c||}{ Theory (BFKL+DL)} \\
\cline{3-6}
$\Delta Y$ & Data --- QPM & \multicolumn{2}{c||}{$\alpha_s[(\tdm k^2+m_q^2)/4]$} &
\multicolumn{2}{c||}{$\alpha_s (\tdm k^2+m_q^2)$} \\
\cline{3-6}
& & Hard & Hard + DL & Hard & Hard + DL \\
\hline\hline
\multicolumn{6}{||c||}{91 GeV} \\ \hline
2 -- 3 & $480 \pm 140 \pm 110$ & 76 & 206 & 34 & 163\\ \hline
3 -- 4 & $240 \pm 60 \pm 50 $ & 114& 237 & 53 & 173\\ \hline
4 -- 6 & $110 \pm 30 \pm 10 $ & 60 & 109 & 29 & 74\\ \hline\hline
\multicolumn{6}{||c||}{183 GeV} \\ \hline
2 -- 3 & $180 \pm 120 \pm 50$ & 51 & 68 & 25 & 42\\ \hline
3 -- 4 & $160 \pm 50 \pm 30 $ & 70 & 86 & 34 & 49\\ \hline
4 -- 6 & $120 \pm 40 \pm 20 $ & 70 & 85 & 35 & 47\\ \hline\hline
\end{tabular}
\end{center}
\end{table}
Using Formula~(\ref{conv}) integrated over the virtualities in the range
allowed by the relevant experimental cuts,
we have calculated the total cross-section for the process
$e^+ e^- \rightarrow e^+ e^- + X$ for LEP1
and LEP2 energies and confronted results of our calculation with the
recent experimental data obtained by the L3 collaboration at LEP \cite{L3}.
Comparison of our results with experimental data is sumarised in Table~1.
We show comparison for $d\sigma/dY$, where $Y=\ln (W^2/Q_1Q_2)$
with subtracted Quark Parton Model (QPM) contribution.
We see that the contamination of the cross-section by soft pomeron
is substantial. The data do also favour the smaller value of the scale of
$\alpha_s$. In general, the results of our calculation lay below the data,
however the error bars are still quite large, so that the discrepancy
is not very pronounced. Let us also mention that cuts applied to obtain
the data shown in Table~1 admit rather low $\gamma\gamma$ energies i.e.
below 10~GeV \cite{L3}, which probably is not sufficient to justify
the validity of high energy limit in QCD.
\section{Exclusive $J/\psi$ production}
The experimental aspects of the measurement of double exclusive
$J/\psi$ production are different from those for the virtual photons
scattering. Namely, since the $c$-quark provides
the energy scale, we may perturbatively describe the cross-section for the
process of exclusive $J/\psi$ production in which almost real photons
take part. It is an important feature beacause the photon flux in electron is
dominated by low virtualities. On the other hand one may measure the
produced $J/\psi$-s through theirs decay products with no need of
tagging of the electrons.
Thus, it is prefered to focus on events with anti-tagged leptons.
The cross-section for the process $e^+e^- \rightarrow e^+e^- + Y$
for anti-tagged $e^{\pm}$ corresponds to the production of the hadronic
state~$Y$ in $\gamma\gamma$ collision and is given by the following
convolution integral: \cite{GGREV}
\begin{equation}
\sigma_{e^+e^- \to e^+e^- + Y} =
\int_0^1 dy_1 \int_0^1 dy_2 \Theta ( W^2 - W^2_{Y0})
\sigma_{\gamma \gamma \rightarrow Y}(W^2)
f_{\gamma/e}(y_1) f_{\gamma/e}(y_2).
\label{conv2}
\end{equation}
where the $\gamma\gamma$ system invariant mass squared $W^2$ is related
to the lepton CM energy squared $s$ by the simple formula: $W^2 = y_1y_2 s$.
The flux factor takes the form:
\begin{equation}
f_{\gamma/e}(y)={\alpha_{em}\over 2 \pi}
\left[ {1 + (1-y)^2 \over y}\;{\rm ln}\,{Q_{max}^2\over Q_{min}^2} -
2 m_e ^2 y \left({1\over Q_{min}^2} - {1\over Q_{max} ^2}\right) \right].
\label{flux}
\end{equation}
and
\begin{equation}
Q_{min}^2 = {m_e^2 y^2 \over ( 1-y)}
\label{pmin}
\end{equation}
\begin{equation}
Q_{max}^2=(1-y)E_{beam}^2 \theta_{max}^2.
\label{pmax}
\end{equation}
The lower limit follows from the kinematics of photon emission from a lepton
whereas the upper one
arises from the upper limit $\theta_{max}$ for the lepton scattering angle.
The minimal invariant mass squared of the hadronic system $W^2_{Y0}$, the
angle $\theta_{max}$ and the beam energy $E_{beam}$ depend on the process
and experimental conditions. For diffractive $J/\psi$ production we shall
choose $\theta_{max}=30$~mrad in accordance with LEP conditions and
$W_{Y0}=15\rm \; GeV$.\\
The formalism that we shall employ to evaluate the cross-section of the
sub-process $\gamma \gamma \rightarrow J/\psi J/\psi$
is very similar to this used in the previous section.
However some modification are neccessary in order to adopt to
specific features of the process.
First of all we have to go beyond the forward configuration of the pomeron
by the use of the BFKL equation with non-zero momentum transver. Besides
that, we introduce a parameter $s_0$ in the propagators of exchanged gluons
instead of the infra-red cut-off $k_0 ^2$ applied in the previous case.
This parameter can be viewed upon as the effective representation of the
inverse of the colour confinement radius squared. Sensitivity of the
cross-section to its magnitude can serve as an estimate of the sensitivity
of the results to the contribution coming from the infrared region. It
should be noted that formula (\ref{ima}) gives finite result in the limit
$s_0=0$. While analyzing this process we use the asymptotic (high-energy)
form of the amplitude, neglecting the phase space effects.\\
The imaginary part ${\rm Im} A(W^2,t=-Q_P ^2)$ of the amplitude for
the considered process which corresponds to the diagram in Fig.~1b
can be written in the following form:
\begin{equation}
{\rm Im} A(W^2,t=- Q_P^2) =
\int {d^2\tdm k\over\pi}{\Phi_0(k^2, Q_P ^2)\Phi(x,\tdm k,\tdm Q_P)\over
[(\tdm k + \tdm Q_P /2)^2 +s_0][(\tdm k - \tdm Q_P /2)^2+s_0]}
\label{ima}
\end{equation}
In this equation $x=m_{J/\psi}^2/W^2$ where $W$ denotes the total
CM energy of the $\gamma \gamma$ system, $m_{J/\psi}$ is the
mass of the $J/\psi$ meson, $\tdm Q_P / 2 \pm \tdm k$ denote the
transverse momenta of the exchanged gluons and $\tdm Q_P$ is the
transverse part of the momentum transfer.
\noindent
\begin{figure}
\epsfxsize = 12cm
\epsfysize = 8cm
\epsfbox{ppfig2.eps}
\caption{\small
The diagrams describing the coupling of two gluons
to the $\gamma \rightarrow J/\psi$ transition vertex.
}
\end{figure}
The impact factor $\Phi_0(k^2, Q_P^2)$ describes the $\gamma J/\psi$
transition induced by two gluons and the diagrams defining
this factor are illustrated in Fig.~3. In the nonrelativistic approximation
they give the following formula for $\Phi_0(k^2, Q_P^2)$
\cite{FORSHAW,GINZBURG}:
\begin{equation}
\Phi_0(k^2, Q_P^2)=
{C\over 2}\sqrt{\alpha_{em}}\alpha_s(\mu^2) \left[{1\over \bar q^2} -
{1\over m_{J/\psi}^2/4+k^2}\right]
\label{impf0}
\end{equation}
where
\begin{equation}
C=q_c{8\over 3} \pi m_{J/\psi} f_{J/\psi}
\label{c}
\end{equation}
with $q_c=2/3$ denoting the charge of a charm quark
and
\begin{equation}
\bar q ^2= {m_{J/\psi}^2+Q_P^2\over 4}
\label{qbar2}
\end{equation}
\begin{equation}
f_{J/\psi}= \sqrt{
{3m_{J/\psi}\Gamma_{J/\psi\rightarrow l^+ l^-}
\over 2\pi \alpha_{em} ^2}
}
\label{fpsi}
\end{equation}
where $\Gamma_{J/\psi \rightarrow l^+ l^-}$ is the leptonic with of the
$J/\psi$ meson. In our calculations we will set
$f_{J/\psi}=0.38{\rm \; GeV}$.
The function $\Phi(x,\tdm k,\tdm Q_P)$ satisfies the non-forward BFKL equation
which in the leading $\ln(1/x)$ approximation has the following form:
$$
\Phi(x,\tdm k,\tdm Q_P)=\Phi_0(k^2, Q_P^2)+ {3\alpha_s(\mu^2)\over
2\pi^2}\int_x^1{dx^{\prime}\over x^{\prime}} \int
{d^2\tdm k' \over (\tdm k' - \tdm k)^2 + s_0} \times
$$
$$
\left\{\left[{{\tdm k_1^2}\over {\tdm k_1^{\prime 2}} + s_0} +
{{\tdm k_2^2}\over {\tdm k_2^{\prime 2}} + s_0}
- Q_P^2
{(\tdm k' - \tdm k)^2+s_0 \over ({\tdm k_1^{\prime 2}} + s_0)
({\tdm k_2^{\prime 2}} + s_0)}
\right]
\Phi(x',\tdm k' ,\tdm Q_P) - \right.
$$
\begin{equation}
\left. \left[{{\tdm k_1^2}\over {\tdm k_1^{\prime 2}} +
(\tdm k' - \tdm k)^2 +2s_0} +
{{\tdm k_2^2}\over {\tdm k_2^{\prime 2}} +
(\tdm k' - \tdm k)^2 +2s_0} \right]
\Phi(x',\tdm k,\tdm Q_P) \right\}
\label{bfkl}
\end{equation}
where
$$
{\tdm k_{1,2}} = {\tdm Q_P \over 2}\pm \tdm k
$$
and
\begin{equation}
{\tdm k_{1,2}^{\prime}} = {\tdm Q_P \over 2} \pm \tdm k^{\prime}
\label{k12}
\end{equation}
denote the transverse momenta of the gluons.
The scale of the QCD coupling $\alpha_s$ which appears in
equations (\ref{impf0}) and (\ref{bfkl}) will be set
$\mu^2=k^2+Q_P^2/4 +m_c^2$ where $m_c$ denotes the mass of the
charmed quark. The differential cross-section is related in the following
way to the amplitude~$A$:
\begin{equation}
{d \sigma \over dt} = {1\over 16 \pi} |A(W^2,t)|^2
\label{dsdt}
\end{equation}
Generalization of the consistency constraint (\ref{cc})
to the case of non-forward configuration
with $Q_P^2 \ge 0$ takes the following form:
\begin{equation}
k'^2 \le (k^2+ Q_P^2/4) {x'\over x}
\label{kc2}
\end{equation}
Besides the BFKL equation (\ref{bfkl}) in the leading
logarithmic approximation we shall also consider the
equation which will embody the constraint (\ref{kc2}) in order to
estimate the effect of the non-leading contributions. \\
The corresponding equation which contains constraint (\ref{kc2}) in the
real emission term reads:
$$
\Phi(x,\tdm k,\tdm Q_P)=\Phi_0(k^2, Q_P^2)+ {3\alpha_s(\mu^2)\over
2\pi^2}\int_x^1{dx^{\prime}\over x^{\prime}}
\int {d^2\tdm k' \over (\tdm k' - \tdm k)^2 + s_0} \times
$$
$$
\left\{\left[{{\tdm k_1^2}\over {\tdm k_1^{\prime 2}} + s_0} +
{{\tdm k_2^2}\over {\tdm k_2^{\prime 2}} + s_0}
- Q_P^2
{(\tdm k' - \tdm k)^2+s_0 \over ({\tdm k_1^{\prime 2}} + s_0)
({\tdm k_2^{\prime 2}} + s_0)}
\right]
\Theta \left((k^2+Q_P^2/4)x'/x-k^{\prime 2}) \right) \times \right.
$$
\begin{equation}
\left.
\Phi(x',\tdm k',\tdm Q_P) -
\left[{{\tdm k_1^2}\over {\tdm k_1^{\prime 2}} +
(\tdm k' - \tdm k)^2 +2s_0} +
{{\tdm k_2^2}\over {\tdm k_2^{\prime 2}} +
(\tdm k' - \tdm k)^2 +2s_0} \right]
\Phi(x',\tdm k,\tdm Q_P) \right\}
\label{bfklkc}
\end{equation}
We solved equations (\ref{bfkl}) and (\ref{bfklkc}) numerically setting
$m_c=m_{J/\psi} /2$.
Brief summary of the numerical method and of the adopted
approximations in solving
equations (\ref{bfkl},\ref{bfklkc}) has been given in Ref.\cite{KMPSI}.
Let us recall that we used running coupling with the scale
$\mu^2=k^2+Q_P^2/4+m_c^2$. The parameter $s_0$ was varied within the range
$0.04 {\rm \; GeV}^2 < s_0<0.16 {\rm \; GeV}^2$. It should be noted
that the solutions of
equations (\ref{bfkl}, \ref{bfklkc}) and the amplitude (\ref{ima}) are
finite in the limit $s_0=0$. This follows from the fact that both impact
factors $\Phi_0(k^2, Q_P^2)$ and $\Phi(x,\tdm k,\tdm Q_P)$ vanish for
$\tdm k=\pm \tdm Q_P/2$ (see equations (\ref{impf0}, \ref{bfkl},
\ref{bfklkc})). The results with finite $s_0$ are however more realistic.
\noindent
\begin{figure}[hbpt]
\leavevmode
\begin{center}
\epsfxsize = 13cm
\epsfysize = 13cm
\epsfbox[18 200 565 755]{ppfig3.ps}
\end{center}
\caption{\small
Energy dependence of the cross-section for the process
$\gamma\gamma \rightarrow J/\psi J/\psi$. The two lower curves
correspond to the calculations based on equation
(\ref{bfklkc}) which contains the non-leading effects coming
from the constraint (\ref{kc2}). The continuous line
corresponds to $s_0=0.04 {\rm \; GeV}^2$ and the dashed line to $s_0=0.16
{\rm \; GeV}^2$. The two upper curves correspond to equation
(\ref{bfkl}) i.e.
to the BFKL equation in the leading logarithmic approximation.
The dashed-dotted line corresponds to $s_0=0.04 {\rm \; GeV}^2$
and short dashed line to $s_0=0.16{\rm \; GeV}^2$.
}
\end{figure}
In Fig.~4 we show the cross-section for the process
$\gamma \gamma \rightarrow J/\psi J/\psi$
plotted as the function of the total CM energy $W$.
We show results based on the BFKL equation in the leading logarithmic
approximation as well as those which include the dominant non-leading
effects. The calculations were performed for the two values of the parameter
$s_0$ i.e. $s_0=0.04 {\rm \; GeV}^2$ and $s_0=0.16 {\rm \; GeV}^2$.
In Fig.~5 we show the $t$-dependence of the cross-section calculated for
$s_0 = 0.10 {\rm \; GeV}^2$.
We show in this figure results for two values of the CM energy $W$
($W=50 {\rm \; GeV}$ and $W=125 {\rm \; GeV}$)
obtained from the solution of the BFKL equation with the non-leading effects
taken into account (see Eq.~(\ref{bfklkc})) and confront them with
the Born term which corresponds to the two (elementary) gluon exchange.
The latter is of course independent of the energy $W$.
The values of the energy $W$ were chosen to be in the region which may
be accessible at LEP2.
Let us discuss crucial features of the obtained results:
\begin{enumerate}
\item {\bf Non leading corrections}. We see from Fig.~4 that the effect of
the non-leading contributions is very important and that they significantly
reduce magnitude of the cross-section and slow down its increase with
increasing CM energy $W$.
\item
{\bf Energy dependence}.
The cross-section exhibits approximate $(W^2)^{2\lambda_P}$
dependence. The parameter $\lambda_P$, which slowly varies with
the energy~$W$ takes the values $\lambda_P \sim 0.23 - 0.28$ within the
energy range $20{\rm \; GeV} < W < 500{\rm \; GeV}$ relevant for LEP2
and for possible TESLA measurements. These results correspond to the
solution of the BFKL equation (\ref{bfklkc}) which contains the non-leading
effects generated by the constraint (\ref{kc2}).
The (predicted) energy dependence of the cross-section
($(W^2)^{2\lambda_P}, \lambda_P \sim 0.23 - 0.28$)
is marginally steeper than that observed in
$J/\psi$ photo-production \cite{VMPHOTOP}. It should however be remebered
that the non-leading effects which we have taken into account although
being the dominant ones still do not exhaust all next-to-leading QCD
corrections to the BFKL kernel \cite{BFKLNL}. The remaining contributions
are expected to reduce the parameter $\lambda_P$ but their effect may be
expected to be less important than that generated by the constraint
(\ref{kc2}). The cross-section calculated from the BFKL equation in the
leading logarithmic approximation gives much stronger energy dependence of
the cross-section (see Fig.~4).
\item
{\bf The value of the cross-section}.
Enhancement of the cross-section is still appreciable after
including the dominant non-leading contribution which follows from the
constraint (\ref{kc2}). Thus while in the Born approximation
(i.e. for the elementary two gluon exchange
which gives energy independent cross-section)
we get $\sigma_{tot} \sim 1.9-2.6$~pb the cross-section calculated from the
solution of the BFKL equation with the non-leading effects taken into account
can reach the value 4~pb at $W=20 \rm \; GeV$ and 26~pb for $W=100 \rm \; GeV$ i.e. for
energies which can be accessible at LEP2.
\item
{\bf Infrared sensitivity}.
The magnitude of the cross-section decreases with increasing
magnitude of the parameter $s_0$ which controls the
contribution coming from the infrared region. This effect is
however much weaker than that generated by the
constraint (\ref{kc2}) which gives the dominant non-leading
contribution. The energy dependence of the
cross-section is practically unaffected by the parameter $s_0$.
\item
{\bf The $t$-dependence}.
Plots shown in Fig.~5 show that the BFKL effects significantly affect
the $t$-dependence of the differential cross-section leading to steeper
$t$-dependence than that generated by the Born term. Possible
energy dependence of the diffractive slope is found to be very
weak (see Fig.~5). Similar result was also found in the BFKL equation
in the leading logarithmic approximation \cite{BARTLQ}.
\end{enumerate}
\begin{figure}
\leavevmode
\begin{center}
\epsfxsize = 13cm
\epsfysize = 13cm
\epsfbox{ppfig4.ps}
\end{center}
\caption{\small
The differential cross-section of the process $\gamma
\gamma \rightarrow J/\psi J/\psi$ corresponding to the solution
of equation (\ref{bfklkc}) which contains the non-leading effects coming
from the consistency (kinematical) constraint (\ref{kc2}) shown for two
values of the CM energy $W$, $W=50{\rm \; GeV}$ (continuous line) and
$W=125 {\rm \; GeV}$ (dashed line). The short dashed line corresponds
to the Born term i.e. to the elementary two gluon exchange
mechanism which gives the energy independent cross-section.
The parameter $s_0$ was set equal to $0.10{\rm \; GeV}^2$.
}
\end{figure}
In our calculations we have assumed dominance of the imaginary part of the
production amplitude. The effect of the real part can be taken into account by
multiplying the cross-section by the correction factor
$1+tg^2(\pi\lambda_P/2)$ which for $\lambda_P \sim 0.25$ can introduce
additional enhancement of about 20~\%.\\
The photonic cross-sections that we obtained in this section are rather
low in terms of the expected number of events, at least for the LEP2
luminosity. Therefore we consider the most inclusive observables relevant for
double $J/\psi$ production in $e^+e^-$ collisions which is the total
cross-section $\sigma_{tot} (e^+ e^- \to e^+ e^- J/\psi J/\psi)$. In fact,
it is convenient to impose additionally the anti-tagging condition.
Taking $\theta_{max}=30$~mrad we get for the
$\sigma_{tot} (e^+ e^- \to e^+ e^- J/\psi J/\psi)$ the values of
about 0.14~pb at $\sqrt{s}=175\rm \; GeV$ and 0.74~pb at $\sqrt{s}=500\rm \; GeV$
(i.e. for typical energies at LEP2 and TESLA respectively). Therefore,
assuming the LEP2 luminosity to be about 500~pb$^{-1}$ we predict about
70~events, which is far below the previous expectations~\cite{GGREV}.
Besides, if one measures both the $J/\psi$-s through the leptonic decay
channels the rate should be divided by factor of about 20, which cuts down
the statistics to only a few events.
\section{Discussion and summary}
From the theoretical point of view, there exist excellent oportunities
to study the exchange of the QCD pomeron in $e^+ e^-$ colliders.
The two golden-plated measurements for this purpose are exclusive $J/\psi$
production and the total $\gamma^* \gamma^*$ cross-section. Both these
processes allow to reduce substantially the contribution of unknown,
nonperturbative elements. However, the leptonic cross-sections in both
cases are well below 1~pb in LEP2 conditions, which makes the measurement
rather difficult there. Nevertheless this problem does not appear at the
future linear colliders $e^+ e^-$ for which the luminosity is expected to
be much larger than at LEP and moreover the cross-section for diffractive
processes is enhanced due to the photon flux and the pomeron effects.
The large expected statistics enables one to reach the region of large
photon virtualities (for double tagged events) where the perturbative
calculations are more reliable. \\
The important point that should be stressed once more is the existence of large
non-leading corrections to BFKL equation, which influence dramatically
the theoretical estimate of the pomeron intercept i.e. the behaviour of the
cross-sections as functions of the energy. The recently calculated magnitude
of next-to-leading contribution to the intercept (for any relevant value of
the strong coupling constant) is comparable or even greater than the leading
term. This implies a very poor convergence of the perturbative series. Thus
one is forced to rely on a resummation scheme.
We adopt the so called consistency
constraint, which is based on the requirement that the
virtualities of gluons exchanged along the ladder are dominated by
transverse momenta squared.
This constraint introduces at the next-to-leading order
a correction to the pomeron intercept which exhausts about 70\% of the
exact QCD result. The main advantage of this approach is that there is a
good physical motivation behind it. Moreover it also offers an approximate
resummation scheme for the perturbative expansion of the intercept.\\
Employing this scheme we found significant reduction of the predicted value of
the intercept in comparison to the leading value.
We find that the calculated behaviour of the
$\gamma^* \gamma^*$ total cross-section exhibits approximate power law
dependence $(W^2)^{\lambda_P}$ with $0.28<\lambda_P<0.35$. It is also found
that the cross-section for $\gamma\gamma \to J/\psi J/\psi$ increases
with increasing energy $W$ as $(W^2)^{2\lambda_P}$ with $\lambda_P$ varying
from~0.23 to~0.28. This has important consequences for the phenomenology,
since the enhancement of the cross-section although
still quite appreciable is much smaller than that which follows from
estimates based on the leading logarithmic approximation \cite{GGREV}.
The results of our calculation are in fair
agreement with the existing data for $\gamma^*\gamma^*$ cross-section from
LEP, although the theoretical calculations have a tendency to underestimate
experimental results.
They are also much more realistic than
the predictions following from the leading order BFKL
equation, which are an order of
magnitude larger. The encouraging element is that even this very first
data with rather low statistics, are enough to show clearly the importance of
non-leading corrections. We may therefore expect that when the excellent data
from linear colliders will be available we will acquire very good opportunity
to test our models and to understand more deeply the physics of the
QCD pomeron. \\
\section*{Acknowledgments}
We are grateful to the Organizers for the interesting and stimulating
Conference. We thank Albert De Roeck for his interest in this work and useful
discussions. This research was partially supported
by the Polish State Committee for Scientific Research (KBN) grants
2~P03B~184~10, 2~P03B~89~13, 2~P03B~084~14 and by the
EU Fourth Framework Programme 'Training and Mobility of Researchers', Network
'Quantum Chromodynamics and the Deep Structure of Elementary Particles',
contract FMRX--CT98--0194.
|
\section{Introduction\protect\\}
\label{sec:intro}
In the absence of a complete formulation of M/string
theory, BPS $p$-brane solutions from various supergravities,
which are the low
energy and/or weak-coupling limits of this intrinsically
non-perturbative unified
theory, are almost the only sources from which reliable non-perturbative
information about this theory can be extracted. These solutions played
vital role in every major development in the past and will continue to do so.
BPS properties obtained in the low energy and weak-coupling limits
remain to be valid
non-perturbatively for the corresponding BPS states.
These properties should also be
independent of whether they are obtained from fields in bulk
spacetime or from fields on
the corresponding worldvolume. For this reason, BPS properties
obtained from the
supergravity $p$-brane
solutions are often used to obtain information about fields living on the
corresponding
worldvolume and vice-versa. For example, the mass per unit $p$-brane
volume calculated from a spacetime BPS Dp-brane solution should
correspond to the Dp-brane
tension of the corresponding Born-Infeld action. This is trivial for simple
Dp-brane
solutions. But it is quite non-trivial for complicated solutions such as BPS
bound state solutions. For example, the tension formula of
$(m, n)$-string bound state of Schwarz \cite{schone}, with $m, n$
relatively prime integers,
can be used to determine the corresponding D-string
worldsheet $U(1)$ field strength, in the linear approximation,
as $g m$ with $g$ the string coupling
\cite{calm}. This in turn can be used to determine the point charge due to
the ending of a fundamental string (for short, F-string) on a general Dp-brane.
After all, a $p$-brane configuration represents the field configuration
created by the corresponding $p$-brane source just like a static electric field
due to a point charge represents the field surrounding the point source. Static
interactions should be the same whether we calculate from the fields or from
the sources. As mentioned above, BPS
properties should be the same no matter whether they are obtained from the
field configuration or from the source. However, apart from these
BPS properties, one must be careful in using them since in general
the fields and
the source are not independent and they actually interact with
each other through the so-called backreaction except in the case where the
so-called decoupling limit is taken. In this limit, the modes propagating on
the worldvolume will decouple from the modes propagating in the bulk spacetime
even though these spacetime modes may have their origins from the worldvolume.
So we expect that in the decoupling limit either the worldvolume fields or
the bulk
fields can be used to describe the same physics. Recently proposed $AdS/CFT$
correspondence by Maldacena \cite{mal} is one such example.
So BPS $p$-brane solutions can provide not only the non-perturbative
information but also a basis for new types of strong-weak dualities. These
are certainly important for us to seek the eventual formulation of M/string
theory, along with the powerful D-brane picture of Polchinski\cite{pol}.
In this paper, we will construct new spacetime solutions for
BPS bound states of Dp-branes carrying certain units of the
quantized worldvolume
constant electric field strength.
We argued in our previous
paper \cite{lurone} that such BPS bound states should exist for
$ 1 \le p \le 8$ in both type IIA (when $p$ is even) and type IIB
(when $p $ is odd) theories\footnote{
In this paper, we specialize in cases of $2 \le p \le 7$.}. The existence of
these bound states was also discussed in \cite{arfs} but in a different approach
of mixed boundary conditions.
The non-vanishing $U(1)$ field
strength on the worldvolume indicates that such a bound state
carries information
about F-strings ending on the Dp-brane, yet preserving one half
of the spacetime
supersymmetries. These Dp-brane bound states should be identified
with the so-called
(F, Dp) bound states which can be obtained from the $(m,n)$-string of
Schwarz or (F, D1) bound state
in type IIB string theory by T-dualities along the transverse directions.
Here integers
$m, n$ are relatively prime. This is precisely the method which
we will use here to construct
these solutions which was also mentioned in \cite{rust,cosp}.
The $p = 3, 4, 6$ configurations for (F, Dp) have been given
in \cite{rust,grelpt,cosp}, respectively. Other non-threshold bound states for
one $p'$-brane within another p-brane with $p' < p$ in M and Type II
string theories have been discussed in \cite{rust,grelpt,cosp,col}.
Our worldvolume picture of these Dp-brane bound states
indicates clearly that the notation `F' in (F, Dp) represents actually
an infinite number of parallel NS-strings.
Precisely, we have
one NS-string per $(2\pi)^{p - 1} \alpha'^{(p - 1)/2}$ area
over a $(p - 1)$-dimensional plane which is perpendicular to
this string.
Each NS-string is $m$ F-strings if the quantized worldvolume
constant field strength
$F_{01} = g m$, with $g$ the string coupling constant (where the field
lines are chosen along
$x^1$ axis). These were obtained in \cite{lurone} by noting that
the (F, D1) worldvolume action can be obtained from a (F, Dp), for
$ p \ge 2$, worldvolume
action by T-dualities. By this, we also determined the tensions for
(F, Dp) bound states
in \cite{lurone}
which clearly indicates that (F, Dp) is a non-threshold bound state.
We will show in
this paper from the corresponding
spacetime solutions that indeed these are all true.
Since we have constructed in this paper the space-time configurations
for all the (F, Dp) bound states in type IIA and IIB theories, it would
be natural to look at how they affect the $AdS/CFT$ correspondences
conjectured by Maldacena. By examining the (F, D3) configuration carefully,
we find that the decoupling limit in this case does not automatically
imply the space-time geometry to be $AdS_5 \times S^5$ as happens in the
case of Maldacena, i.e., for simple D3-branes. We also find that the string
coupling constant in this case is replaced by a finite but a smaller
effective string coupling constant. However, we observed that by a suitable
rescaling of the coordinates the $AdS_5/CFT_4$ correspondence may still hold
true, but now with respect to (F, D3) bound state rather than the simple
D3-branes. The effective string coupling constant in the near horizon region
is quantized in terms of the integer $n$, the 5-form flux and the integer
$m$, the number of F strings per $(2\pi)^2 \alpha'$ area over the two
dimensional plane perpendicular to the F strings, where $m$, $n$ are
relatively prime integers. This coupling can be independent of the usual
string coupling at spatial infinity in the limit $m \gg n$.
This paper is organized as follows. In the following section,
we discuss in
some detail
the so-called vertical and diagonal (or double) dimensional
reductions/oxidations
in type
IIA and type IIB supergravities which are the basic methods we use
to construct
the (F, Dp) configurations. In section 3, we first give a detail
construction of
(F, D2) as an illustrating example. We then list results for each of
(F, Dp) for
$ 3 \le p \le 7$. We also present the (W, Dp) (with W representing the waves)
solutions for $0 \le p \le 6$
from these newly found (F, Dp) configurations by T-dualities along
the direction
of F-strings. In section 4, we calculate the charge per unit
$(p - 1)$-dimensional area for the F-strings and show that we indeed have
$m$ F-strings
per $(2\pi)^{p -1} \alpha'^{(p - 1)/2}$ of ($p - 1$)-dimensional area for
these bound states.
We also calculate the mass per unit $p$-brane volume for these
solutions and show that the
corresponding tensions agree precisely with what we obtained in
\cite{lurone} based on
the worldvolume study. In section 5, we study the decoupling limit for the
(F, D3) bound state. Finally, we conclude this paper in section 6.
\section{Diagonal and Vertical Dimensional Reductions/Oxidations\protect\\}
\label{sec:dvdr}
It is well-known that type IIA and type IIB superstring theories
compactified on a circle are equivalent by T-duality i.e.
order by order in perturbation theory. More precisely,
IIA and IIB theories are interchanged by the T-duality transformation
$R \rightarrow 1/R$, with $R$ the compactified radius, along with
an interchange of
momentum or KK modes with the winding modes. However, this equivalence
does not extend to the
respective $S^1$-compactified supergravity theories due to the absence of the
string winding modes. But this will not have any effect on the massless
modes in $D = 9$.
Therefore, the $D = 9, N = 2$ supergravity obtained from type IIA supergravity
by the
dimensional reduction must be equivalent by T-duality to that obtained
from type IIB supergravity. This also
follows from supersymmetry because $D = 9, N = 2$ supergravity is unique
up to field
redefinitions. Hence, the T-duality acting on the $D = 9$ massless
fields must map
fields in IIA basis to those in IIB basis. This feature, as we will see,
is precisely what we
need to obtain a BPS solution in type IIA theory from a BPS solution in
type IIB theory and vice versa by
a T-duality transformation.
It is well-known that type II theories contain the following simple
BPS states\footnote{
By simple, we mean an object whose charge is associated with only
one antisymmetric
tensor field in
type II theories. We have not included here the D9 branes, i.e., the
so-called spacetime-filling branes\cite{ber}.}:
\begin{eqnarray}
{\rm Type~ IIA}~\qquad&& {\rm W}~~{\rm F}~~{\rm NS5}~~{\rm KK}~~
\qquad~{\rm D0}~~{\rm D2}~~{\rm D4}~~{\rm D6}~~{\rm D8}\nonumber\\
{\rm Type~ IIB}~\qquad&& {\rm W}~~{\rm F}~~{\rm NS5}~~{\rm KK}~
\qquad~{\rm D(-1)}~~{\rm D1}~~{\rm D3}~~{\rm D5}~~{\rm D7}\nonumber\\
\end{eqnarray}
where W, F, NS5 and KK denote waves, fundamental strings, NS fivebranes,
and KK
monopoles, respectively, and they are associated with the NSNS fields.
Also, Dp ($ - 1\le p \le 8$)
are the so-called D-branes and they are associated with the so-called
RR fields.
The action of T-duality along parallel or transverse direction
on the above objects present in type II theories
is given in the following table.
\begin{center}
\begin{tabular}{|c|c|c|} \hline
& Parallel
& Transverse\\
\hline
Dp& D(p $-$ 1) & D(p + 1)\\
F& W & F \\
W& F & W \\
NS5& NS5 &KK \\
KK & KK & NS5 \\
\hline\end{tabular}
\end{center}
The T-duality may be performed along either one of the
worldvolume directions or one of the transverse directions
(for KK monopole, the transverse
direction is taken to be the nut direction).
It is clear from the above table that if we start from a (F, D1) bound
state in
the type IIB theory, we will have (F, D2) bound state in type IIA theory by
T-dualizing (F, D1) along one of its transverse directions. Then we can have
(F, D3) bound state in the type IIB theory by T-dualizing (F, D2) along one of
its transverse directions. Repeating this procedure, we can have (F, D4) from
(F, D3), (F, D5) from (F, D4), (F, D6) from (F, D5) and finally (F, D7) from
(F, D6). We cannot obtain (F, D8) from (F, D7) by this procedure. We will
explain possible reasons behind this at the end of this section.
If we T-dualize
the above bound
states along the longitudinal direction of F in (F, Dp) for $ 1 \le p \le 7$,
we have (W, Dp) bound states for $ 0 \le p \le 6$ with W
representing the waves.
The question now is how to implement the above T-duality
at the level of solutions of the IIA and IIB supergravities.
As discussed above,
the $S^1$ compactified IIA and IIB supergravities are not equivalent by
T-duality except for the massless sector of $D = 9, N = 2$ supergravity.
This implies that
T-duality can be implemented in obtaining solutions in one supergravity
from the known solutions each with one isometry along the
would-be compactified direction in the other supergravity. This is exactly the
rationale behind the procedure of the so-called vertical dimensional reduction
and diagonal (or double) dimensional oxidation or the diagonal reduction and
vertical oxidation, depending on whether the T-duality acts along a transverse
direction or a worldvolume direction.
In this paper, we choose a convention such that $H_n$ denotes an
$n$-form field
strength in type IIB supergravity or in $D = 9, N = 2$ supergravity
in terms of the
type IIB basis
and similarly $F_n$ denotes an $n$-form in type IIA supergravity or in
$D = 9, N = 2$
supergravity
in terms of the type IIA basis. If an $n$-form is reduced to
an $(n-1)$-form
from $D = 10$ to $D = 9$, we denote the resulting $(n-1)$-from with
a superscript `1'
indicating that it is obtained by one-step reduction.
For example, if the reduction is
along one of the indices of $H_n$, we have
$H_n \rightarrow H^{(1)}_{n - 1} \wedge dz$.
In the case of T-duality along a transverse
direction of a BPS $p$-brane solution ($ p \le 6$) as discussed in
\cite{dabghw,lups}, we
first
use the ``no-force" condition
for the $p$-brane configuration to obtain a multi-center solution from a
single center
one by placing many $p$-branes parallel to each other. Without loss
of generality and to be
specific, let us assume that this $p$-brane is in type IIB supergravity.
The multi-center solution can be obtained by
the following replacement in the corresponding
Harmonic function. For example, if a single-center Harmonic function
$H = 1 + Q_p /r^{7 - p}$ with $r^2 = (y^1)^2 + (y^2)^2 + \cdots +
(y^{9 - p})^2$, then the multi-center Harmonic function will be given as,
\begin{equation}
H = 1 + \sum_n \frac{Q_p^{(n)}}{\mid \vec{y} - \vec{y}_n \mid^{7 - p}}.
\label{eq:mcs}
\end{equation}
In order to generate one isometry along the T-dual direction, say,
$y^{9 - p}$-direction, as pointed out in \cite{lups}, a continuous
stack of $p$-branes with uniform charge density should be placed along the
coordinate $z = y^{9 - p}$ such that the summation in
Eq.\ (\ref{eq:mcs}) can be replaced by an integration\footnote{This seems
to work also for non-BPS solutions as discussed in \cite{lups}.}. For
the purpose
of this paper, let us take a close look at this continuum limit. We first
take $\vec{y}_n = 2 \pi n a \hat{z}$ in Eq.\ (\ref{eq:mcs}),
with $n \in Z$,
$\hat{z}$ the unit
vector along $y^{9 - p}$-direction and $a = \sqrt{\alpha'}$ where $\alpha'$
is the string constant. If we allow the
range of the summation in Eq.\ (\ref{eq:mcs}) to be
$-\infty < n < \infty$ and
take all charges to be equal, i.e., $Q_p^{(n)} = Q_p$, then the
multi-center solution
is an infinite periodic array of $p$-branes. The solution would be a
stringy one if we take
$\alpha' \rightarrow 0$. Under this limit, the summation
in Eq.\ (\ref{eq:mcs}) can also
be replaced by an integration, i.e.,
\begin{equation}
Q_p \sum_{n = - \infty}^\infty \frac{1}{\mid \vec{y} - 2\pi n a
\hat{z} \mid^{7 - p}}
\rightarrow \frac{Q_p}{2\pi a} \int_{- \infty}^\infty \frac{dz}
{(r^2 + z^2)^{(7 - p)/2}}
= \frac{\tilde{Q}_p}{ r^{6 - p}},
\end{equation}
where $\tilde{Q}_p = c_p Q_p$ with constant $c_p = Q_{p + 1}/Q_p$.
The explicit
expressions for $Q_{p + 1}$ and $Q_p$ are given in the next section.
In the above,
we have now $r^2 = (y^1)^2 + \cdots + (y^{8 - p})^2$. In other
words, one isometry is generated along the T-dual direction by the
infinite periodic array
of $p$-branes. Once this solution is known, a BPS $p$-brane solution in
$D = 9$ can be obtained from
it. The Einstein metric for $D = 9$ $p$-brane can be obtained from
the corresponding metric of
$D = 10$ $p$-brane as (see, for example, the appendix of \cite{duflp})
\begin{equation}
ds^2_{10}~ ({\rm type~ IIB})= e^{-\varphi_B/(2\sqrt{7})} ds_9^2 +
e^{\sqrt{7} \varphi_B/2} d z^2,
\label{eq:mriib}
\end{equation}
where $\varphi_B$ is a dilatonic scalar originating from the
dimensional reduction.
Since we know $d s^2_{10}~({\rm type~ IIB})$ and it also has
one isometry along
$z$-direction,
we can read from
the above
the new field $\varphi_B$ and therefore the $D = 9$ Einstein metric
$d s_9^2$. The $D = 10$ dilaton
$\phi_B$ remains the
same as the one for the solution of the infinite periodic array of
$p$-branes in $D = 10$. The rule for reducing
an $n$-form
field strength $H_n$ is as follows: if it carries an electric-like charge,
it will remain the same
in $D = 9$
whereas it will reduce to an $(n-1)$-form, i.e.,
$H_n \rightarrow H^{(1)}_{n - 1}\wedge dz$,
if it carries a magnetic charge. So we now obtain a $D = 9$ BPS
$p$-brane solution from a $D = 10$ BPS $p$-brane solution by the
so-called vertical dimensional reduction.
We expect that the T-dual of the above $p$-brane in type IIB gives a
$(p + 1)$-brane in type IIA.
Let us first pretend that we know this $(p + 1)$-brane solution in type IIA.
We denote the brane coordinate
$x^{ p + 1} = z$ ( We choose the brane along $x^1, \cdots,
x^{p + 1}$-directions). All the fields for a
static BPS $(p + 1)$-brane are usually independent of the brane coordinates.
So we have one isometry already
along the $z$-direction. Thus a BPS $p$-brane solution in $D = 9$
can be obtained simply by
the so-called diagonal or double dimensional reduction \cite{dufhis} on the
BPS $(p + 1)$-brane solution of type IIA supergravity.
The two Einstein metrics are now simply related to
each other by
\begin{equation}
d s^2_{10}~ ({\rm type~ IIA}) = e^{-\varphi_A/(2\sqrt{7})}ds_9^2
+ e^{\sqrt{7} \varphi_A/2} d z^2,
\label{eq:mriia}
\end{equation}
where $\varphi_A$ is another dilatonic scalar due to this
dimensional reduction.
The IIA dilaton $\phi_A$ remains unchanged. The $n$-form field strength
$F_n$ is reduced according to
the following:
if it carries an electric charge, it will become an $(n-1)$-form in
$D = 9$, i.e.,
$F_n \rightarrow F^{(1)}_{n -1} \wedge dz$, whereas it will remain the
same if
it carries a magnetic charge, just the opposite to the case of
vertical reduction. Here we do not really
change anything. It is just the same whether we call the solution a
$(p + 1)$-brane in $D = 10$ or a $p$-brane
in $D = 9$. In other words, if we know a $p$-brane solution in $D = 9$,
we should obtain a $(p + 1)$-brane solution in
$D = 10$ right away and vice-versa. The process of obtaining a
$(p + 1)$-brane in $D = 10$ from a $p$-brane
in $D = 9$ is called the diagonal oxidation.
The $p$-brane solution in $D = 9$ obtained from a known type IIB $p$-brane
solution is described in terms of
fields in type IIB basis while in obtaining a $(p + 1)$-brane in type IIA,
we need a $p$-brane solution in $D = 9$
described in terms of the fields in type IIA basis. So we need to map
the fields in type IIB basis to fields
in type IIA basis. This mapping is nothing but the T-duality transformation
for the solution. This is also
the T-duality transformation acting on the fields of $D = 9, N = 2$
supergravity
discussed earlier. Then T-duality on field strengths in $D = 9, N = 2$
supergravity or the corresponding $p$-brane solutions is the identification
$H_n = F_n$ with the understanding that if $H_n$ is defined with
a Chern-Simons term
so is $F_n$. The mapping for the $D = 10$ dilaton $\phi$ and the dilatonic
scalar $\varphi$ between the two versions of $D = 9, N = 2$ supergravity is
\begin{equation}
\left(\begin{array}{c}
\phi_A\\
\varphi_A \end{array}\right)
= \left(\begin{array}{cc}
\frac{3}{4}& - \frac{\sqrt{7}}{4}\\
- \frac{\sqrt{7}}{4} & - \frac{3}{4} \end{array}\right)
\left(\begin{array}{c}
\phi_B\\
\varphi_B \end{array}\right).
\label{eq:tdvdr}
\end{equation}
The above relations have been given or can be deduced, for example, from
the appendix in \cite{duflp} where the relations between the corresponding
gauge
potentials are also given\footnote{The precise relations between various fields in two different
bases of the nine dimensional theory are given in a recent
paper \cite{lupsone}. We would like to thank Chris Pope for letting
us know their results prior to publication. These relations are also given in
\cite{berho} but in a different conventions than we use here.}.
Using the above relations, we can obtain a $p$-brane solution in $D = 9$
in type IIA basis from a $p$-brane solution in type IIB basis.
Then we obtain
a BPS $(p + 1)$-brane solution in type IIA by oxidizing
this $p$-brane solution back to $D = 10$ as described above.
If we reverse the above process, we then obtain a BPS $p$-brane solution
in one
theory from a known BPS $(p + 1)$-brane solution in the other theory now by
a T-duality along a world volume direction. The oxidation involved in this
process is the so-called vertical one\cite{lups}.
Now we conclude this section by explaining why we cannot
obtain D8 brane
solution in the massive type IIA supergravity from a D7 solution in the type
IIB supergravity by a T-duality transformation along one of its transverse
directions. First, it is not clear even now whether the massive type IIA
supergravity has its origin in any of the known superstring theories or
even in M-theory (though we believe so). So it is not clear that there exists
a theory from which the massive type IIA supergravity originates and is T-dual
to the type IIB superstring theory. Even if we have such a theory,
the massless
fields in $D = 9$ from the $S^1$-compactified massive type IIA supergravity
cannot describe the 8-brane in $D = 9$. Therefore, we do not expect that the
D8-brane solution is related to a D7 brane solution in the type IIB
supergravity
by a T-duality transformation\footnote{We merely mean this at the level of
supergravity solutions. As we know, in string theory, D8 brane does seem to
be related to D7 brane by
a T-duality.}. Technically, as explained in \cite{lups}, the
vertical reduction does not seem to apply when $p = 7$. By the same token, we
cannot obtain (F, D8) solution from (F, D7) solution by a T-duality
transformation.
\section{(F, D$_p$) Bound States\protect\\}
\label{sec:fdpbs}
In this section, we will construct explicitly the (F, Dp) bound state
configurations for $ 2 \le p \le 7$ from the known (F, D1) bound state
configuration of Schwarz \cite{schone} by a T-duality transformation
along one of the transverse directions described in the previous section.
We will give a detail construction for the case of (F, D2) as an example and
present only the results for the rest. Once we have (F, Dp) configurations for
$ 1\le p \le 7$, we will also present the results for (W, Dp) for
$0 \le p \le 6$, i.e., a Dp brane
carrying waves in it, by a T-duality along the longitudinal direction
of F-strings.
All these configurations preserve one half of the spacetime supersymmetries.
The field configurations of the $(m,n)$-string of Schwarz
(or (F, D1) bound state), with
$m, n$ relatively prime integers, in type IIB theory \cite{schone}
are given in terms of our notations as: the Einstein metric,
\begin{equation}
ds^2~ ({\rm type ~IIB}) = H^{-3/4} [ - (d x^0)^2 + (d x^1)^2 ]
+ H^{1/4} dy^i dy^i,
\label{eq:fd1m}
\end{equation}
with $i = 1, \cdots, 8$; the type IIB dilaton,
\begin{equation}
e^{\phi_B} = e^{\phi_{B0}}\, H' H^{-1/2},
\label{eq:iibd}
\end{equation}
the axion,
\begin{equation}
\chi_B = \frac{mn (H - 1) + \chi_{B0} \,\Delta_{(m, n)} \,e^{\phi_{B0}}}
{ n^2 H + (m - \chi_{B0} n)^2 \,e^{2\phi_{B0}}},
\label{eq:iiba}
\end{equation}
and the NSNS 3-form field strength $H_3 ~({\rm NSNS})$ and
RR 3-form $H_3~({\rm RR})$,
\begin{eqnarray}
H_3~ ({\rm NSNS}) &=& - \Delta_{(m, n)}^{-1/2} \,e^{\phi_{B0}}
\,(m - \chi_{B0} n) d H^{-1}
\wedge d x^0 \wedge d x^1,\nonumber\\
H_3~ ({\rm RR})& = &\Delta_{(m, n)}^{-1/2} \left[ \chi_{B0}\,
(m - \chi_{B0} n)\,
e^{\phi_{B0}} - n \,e^{-\phi_{B0}}\right] d H^{-1}\wedge d x^0 \wedge d x^1.
\label{eq:3f}
\end{eqnarray}
In the above, $\phi_{B0}$ and $\chi_{B0}$ represent the asymptotic
values of the
type IIB dilaton and axion, respectively. The SL(2, Z) invariant
$\Delta$-factor is
\begin{equation}
\Delta_{(m, n)} = (m - \chi_{B0} n)^2 e^{\phi_{B0}} + n^2 e^{-\phi_{B0}},
\label{eq:df}
\end{equation}
and the SL(2, Z) invariant Harmonic function $H$ is
\begin{equation}
H = 1 + \frac{Q_1}{r^6},
\label{eq:hf1}
\end{equation}
where the radial distance $r^2 = (y^1)^2 + (y^2)^2 + \cdots + (y^8)^2$ and
the quantized central charge $Q_1$ is
\begin{equation}
Q_1 = 2^5 \pi^2 \alpha'^3 \Delta_{(m, n)}^{1/2},
\label{eq:cc1}
\end{equation}
with $\alpha'$ the string constant. We have also introduced a second
Harmonic function
\begin{equation}
H' = \frac{ (m - \chi_{B0} n)^2 e^{\phi_{B0}} + n^2 H e^{-\phi_{B0}}}
{\Delta_{(m, n)}},
\label{eq:nhf2}
\end{equation}
which approaches unity as $r \rightarrow \infty$.
Before we move on to the constructions of (F, Dp) bound states,
we fix a few
conventions. The RR-charge of the Dp-brane in (F, Dp) is defined
\cite{dufkl} as,
\begin{equation}
e_p = \frac{1}{\sqrt{2} \kappa_0} \int_{S^{8 - p}} e^{ - a(p)
\phi} \ast G_{p
+2},
\label{eq:ecd}
\end{equation}
for Noether ``electric" charge, and
\begin{equation}
g_p = \frac{1}{\sqrt{2} \kappa_0} \int_{S^{p + 2}} G_{p + 2},
\label{eq:mcd}
\end{equation}
for topological or magnetic-like charge. In the above $\sqrt{2} \kappa_0 =
(2\pi)^{7/2} \alpha'^2$. For a Noether charge,
the integration is over an asymptotic $(8 - p)$-sphere surrounding the
Dp-brane
while for a magnetic-like charge, the integration is over an asymptotic
$(p + 2)$-sphere surrounding the Dp-brane. Also, $G_{p + 2} = H_{p + 2}$
for type IIB
theory while $G_{p + 2} = F_{p + 2}$ for type IIA theory.
The constant $a (p) = (p - 3)/2$ for Dp-branes in both type IIA
and type IIB theories. The above definitions are also valid for NSNS branes,
i.e., NSNS strings and fivebranes, but with $a (p) = - (p - 3)/2$.
However, the
charge associated with the F-strings in the (F, Dp) bound states cannot be
calculated using the above simple formula. We will discuss this in the next
section. In this paper, $\ast$ always denotes the Hodge dual. $\epsilon_n$
denotes the volume form on an $n$-sphere where the volume of
a unit $n$-sphere is
\begin{equation}
\Omega_n = \frac{2 \pi^{(n + 1)/2}}{\Gamma ((n + 1)/2)},
\label{eq:usv}
\end{equation}
and the unit charge
for a Dp-brane is given as,
\begin{equation}
Q_0^p \equiv (2\pi)^{(7 - 2p)/2} \alpha'^{(3 - p)/2}.
\label{eq:puc}
\end{equation}
{\bf (F, D2) Bound State}: Now in order to construct this bound state we
first need to construct, as described in the previous section, the solution
corresponding to an infinite periodic array of $(m,n)$-strings along
$z = y^8$ axis.
The Harmonic function in that case would be given as,
\begin{equation}
H = 1 + Q_1 \sum_{n = -\infty}^\infty \frac{1}{\mid \vec{y}
- 2\pi n a \hat{z}\mid^6},
\label{eq:mcs1}
\end{equation}
where $\hat{z}$ denotes the unit vector and $a = \sqrt{\alpha'}$.
In the limit $\alpha' \rightarrow 0$, the
summation in the above Harmonic function can be replaced by an
integration, i.e.,
\begin{equation}
\sum_{n = -\infty}^\infty \rightarrow \int_{-\infty}^\infty \frac{dz}{2\pi a}.
\label{eq:sir}
\end{equation}
Note that writing $\vec{y} = \tilde{\vec{y}} + z \hat{z}$ and so,
$\mid \vec{y} - 2\pi na \hat{z} \mid^6 = [ {\tilde{r}}^2
+ (z - 2\pi na)^2]^3$, with
${\tilde r}^2 = (y^1)^2 + \cdots + (y^7)^2$, we have,
\begin{equation}
\sum_{n = -\infty}^\infty \frac{1}{\mid \vec{y} - 2
\pi n a \hat{z}\mid^6}
\rightarrow \int_{-\infty}^\infty \frac{dz}{2\pi a} \frac{1}
{({\tilde r}^2 +
z^2)^3} = \frac{3}{16 a} \frac{1}{{\tilde r}^5}.
\label{eq:ir1}
\end{equation}
So the new Harmonic function describing the infinite array of $(m,n)$-strings
along the $z$-direction is
\begin{equation}
H = 1 + \frac{Q_2}{r^5},
\label{eq:nh2}
\end{equation}
where we have dropped the `tilde' above the radial distance and
will continue to do so
from now on. The
central charge $Q_2$ in (3.16) is
\begin{equation}
Q_2 = \frac{3 Q_1}{16 \alpha'^{1/2}} = 6 \pi^2 \alpha'^{5/2}
\Delta_{(m, n)}^{1/2},
\label{eq:cc2}
\end{equation}
The Harmonic function is now independent of the $z$-coordinate.
In other words,
the Einstein metric (Eq.\ (\ref{eq:fd1m})) expressed in terms of this
Harmonic
function possesses one additional isometry along the $z$-direction. If we now
compactify the metric along $z$-direction, Eq.\ (\ref{eq:mriib}) will give
\begin{equation}
e^{\sqrt{7} \varphi_B /2} = H^{1/4},
\label{eq:vd2}
\end{equation}
and the 9-d Einstein metric will be given as,
\begin{equation}
d s^2_9 = H^{- 5/7} [ - (d x^0)^2 + (d x^1)^2 ] + H^{2/7} dy^i dy^i,
\label{eq:m9}
\end{equation}
where $i = 1, 2, \cdots, 7$. For the 9-dimensional solution, the type IIB
dilaton, axion and NSNS and RR 3-forms remain unchanged with the
replacement of the
old Harmonic function by the new one. We have therefore obtained
$(m,n)$-string bound
state in $D = 9$ in the type IIB basis.
We now express all the relevant fields in type IIA basis. The $d s_9^2$
remains
unchanged. $F_3 = H_3 ~({\rm NSNS})$, $F_3^{(1)} = H_3 ~({\rm RR})$ and
$F_1^{(1)} = d \chi$. The type IIA dilaton $\phi_A$ and the dilatonic scalar
$\varphi_A$ can be obtained from $\phi_B$ and $\varphi_B$ through
Eq.\ (\ref{eq:tdvdr}) as
\begin{eqnarray}
e^{\phi_A} &=& e^{3\phi_{B0}/4} H'^{3/4} H^{- 1/2},\nonumber\\
e^{\sqrt{7} \varphi_A} &=& e^{-7 \phi_{B0}/4} H'^{- 7/4} H^{1/2},
\label{eq:iiadvd}
\end{eqnarray}
where the Harmonic function $H'$ continues to be given by the expression
\ (\ref{eq:nhf2}) but in terms of the present Harmonic function $H$.
Once we express $D = 9$ (F, D1) configuration in the type IIA
basis, we can read the (F, D2) configuration in type IIA theory using the
diagonal oxidation described in the previous section. For example, the
Einstein metric in $D = 10$ for this configuration can be read from
Eq.\ (\ref{eq:mriia}) with $d s^2_9$, $\phi_A$ and
$\varphi_A$ given in (3.19) and (3.20). We here collect the complete
results for the (F, D2)
configuration for the metric,
\begin{equation}
d s^2_{10}~ ({\rm type~ IIA}) = e^{\phi_{B0}/8}
H'^{1/8} H^{1/4} \left[ H^{-1}
(- (d x^0)^2 + (d x^1)^2 ) + e^{ - \phi_{B0}} H'^{ - 1} (dx^2)^2
+ d y^i d y^i\right],
\label{eq:iiadtwom}
\end{equation}
for the dilaton,
\begin{equation}
e^{\phi_A} = e^{3\phi_{B0}/4} H'^{3/4} H^{-1/2},
\label{eq:iiadtwod}
\end{equation}
and for the remaining non-vanishing fields\footnote{In type IIA theory,
$F_4' = dA_3 + A_1\wedge F_3$},
\begin{eqnarray}
F_2 &=& n \Delta_{(m,n)}^{-1} ~e^{-\phi_{B0}}
(m -\chi_{B0} n)~ H'^{-2} ~d H\wedge
dx^2,\nonumber\\
F_3 &=& - \Delta_{(m,n)}^{-1/2} ~e^{\phi_{B0}} ~(m - \chi_{B0} n)
~d H^{-1}\wedge
d x^0\wedge d x^1,\nonumber\\
F_4' &=& n \,e^{- 3 \phi_{B0}/8}\, H'^{-3/8} \, H^{1/4}\,
\frac{\sqrt{2} \kappa_0 Q_0^2}
{\Omega_5} \,\ast \epsilon_6.
\label{eq:tstwo}
\end{eqnarray}
In the above, $i = 1, 2, \cdots
7$, the Harmonic functions $H'$ and $H$ are given by
Eqs.\ (\ref{eq:nhf2}) and (\ref{eq:nh2}), respectively.
{\bf (F, D3) Bound State}\footnote{This configuration with zero
asymptotic values of
$\phi_{B0}$ and $\chi_{B0}$ was also given in
\cite{rust} for different purpose. Some non-threshold bound
states of M theory,
not considered here, were also discussed there.}: Once
we obtain (F, D2), we can repeat the above
process to obtain (F, D3) bound state configuration. The results are for
the Einstein metric,
\begin{eqnarray}
ds^2_{10}~ ({\rm type ~IIB}) = &&e^{\phi_{B0}/4} \,H'^{1/4}\,
H^{1/4} [ H^{-1} (
- (d x^0)^2 + (d x^1)^2) \nonumber\\
&& + e^{-\phi_{B0}}\, H'^{-1} ( (d x^2)^2 + (d x^3)^2)
+ d y^i d y^i],
\label{eq:miibthree}
\end{eqnarray}
with $i = 1, 2, \cdots, 6$; for the type IIB dilaton,
\begin{equation}
e^{\phi_B} = e^{\phi_{B0}/2}\, H'^{1/2}\, H^{-1/2},
\label{eq:iibdthree}
\end{equation}
and for the remaining non-vanishing fields,
\begin{eqnarray}
H_3 ~({\rm NSNS}) &=& - \Delta_{(m,n)}^{-1/2} \,e^{\phi_{B0}}
\,(m - \chi_{B0} n) \,
d H^{-1}\wedge d x^0 \wedge d x^1,\nonumber\\
H_3 ~({\rm RR}) &=& n \,\Delta_{(m,n)}^{-1}\, e^{-\phi_{B0}}
(m - \chi_{B0} n)
H'^{-2 } \,d H \wedge d x^2 \wedge d x^3,\nonumber\\
H_5 &= & n \,\frac{\sqrt{2} \kappa_0 Q_0^3 }{\Omega_5} \,(\ast \epsilon_5 +
\epsilon_5).
\label{eq:iibthreerf}
\end{eqnarray}
In the above, the Harmonic function $H'$ continues to be given by
Eq.\ (\ref{eq:nhf2}) but the Harmonic function $H$ is now
\begin{equation}
H = 1 + \frac{Q_3}{r^4},
\label{eq:hfthree}
\end{equation}
where $Q_3 = \Delta_{(m,n)}^{1/2} \sqrt{2} \kappa_0 Q_0^3
/(4 \Omega_5)$. All the other
quantities have already been defined.
{\bf (F, D4) Bound State}\footnote{The classical configuration
of this bound state
was also given in \cite{grelpt}, obtained by dimensional reductions from
the (M2, M5) bound state in D = 11.}:
Repeating the same procedure the various field
configuration for this solution are:
the metric,
\begin{eqnarray}
d s^2_{10}~ ({\rm type~ IIA}) =&& e^{3 \phi_{B0}/8}\, H'^{3/8} \, H^{1/4}\,
[ H^{-1}\, (- (d x^0)^2 +
(d x^1)^2) \nonumber\\
&& + e^{-\phi_{B0}}\, H'^{-1} \,( (d x^2)^2 + (d x^3)^2 + (d x^4)^2)
+ d y^i d y^i],
\label{eq:iiamfour}
\end{eqnarray}
with $i = 1, 2, \cdots, 5$; the type IIA dilaton,
\begin{equation}
e^{\phi_A} = e^{\phi_{B0}/4}\, H'^{1/4} \,H^{- 1/2},
\label{eq:iiadfour}
\end{equation}
and the remaining non-vanishing fields,
\begin{eqnarray}
F_3 &= &- \Delta_{(m,n)}^{-1/2} \, e^{\phi_{B0}}\,
(m - \chi_{B0} n)\, d H^{-1}
\wedge d x^0 \wedge d x^1,\nonumber\\
F_4' &=& n \frac{\sqrt{2} \kappa_0 Q_0^4}{\Omega_4} \,\epsilon_4 +
n \, \Delta_{(m,n)}^{-1} \,e^{-\phi_{B0}}\, (m
- \chi_{B0} n) \,H'^{-2}\, d H \wedge d x^2 \wedge d x^3 \wedge d x^4.
\label{eq:iiarffour}
\end{eqnarray}
In the above, the Harmonic function $H'$ is as given by Eq.\ (\ref{eq:nhf2})
but the Harmonic
function $H$ is now
\begin{equation}
H = 1 + \frac{Q_4}{r^3},
\label{eq:iiahffour}
\end{equation}
where $Q_4 = \Delta_{(m,n)}^{1/2} \sqrt{2} \kappa_0 Q_0^4 /(3 \Omega_4)$.
{\bf (F, D5) Bound State}: The field configurations for this bound state are:
the Einstein metric,
\begin{eqnarray}
d s^2_{10} ~({\rm type~ IIB}) =&& e^{\phi_{B0}/2}\, H'^{1/2} \, H^{1/4} \,
[ H^{-1} (- (d x^0)^2 +
(d x^1)^2) \nonumber\\
&&+ e^{-\phi_{B0}}\, H'^{-1} ( (d x^2)^2 + (d x^3)^2
+ (d x^4)^2 + (d x^5)^2)
+ d y^i d y^i],
\label{eq:iibmfive}
\end{eqnarray}
with $i = 1, 2, 3, 4$; the type IIB dilaton,
\begin{equation}
e^{\phi_B} = H^{- 1/2},
\label{eq:iibdfive}
\end{equation}
and the remaining non-vanishing fields,
\begin{eqnarray}
H_3 ~({\rm NSNS}) &=& - \Delta_{(m,n)}^{-1/2}\, e^{\phi_{B0}}\,
(m - \chi_{B0} n)\, d H^{-1}\wedge d x^0
\wedge d x^1, \nonumber\\
H_3 ~({\rm RR}) & = & n \,\frac{\sqrt{2} \kappa_0 Q_0^5}{\Omega_3}\,
\epsilon_3,\nonumber\\
H_5 &=& n \,\Delta_{(m,n)}^{-1} \,e^{-\phi_{B0}} \,(m
- \chi_{B0} n)\, H'^{-2} \,d H \wedge d x^2 \wedge d x^3 \wedge d x^4
\wedge d x^5.
\label{eq:iibrffive}
\end{eqnarray}
Again the Harmonic function $H'$ remains the same as given in
Eq.\ (\ref{eq:nhf2}) but the Harmonic
function $H$ is now
\begin{equation}
H = 1 + \frac{Q_5}{r^2},
\label{eq:iibhffive}
\end{equation}
where $Q_5 = \Delta_{(m,n)}^{1/2} \sqrt{2} \kappa_0 Q_0^5 /(2 \Omega_3)$.
{\bf (F, D6) Bound State}\footnote{The classical configuration of this bound state
was also given in \cite{cosp}.}: In this case we have the Einstein metric,
\begin{eqnarray}
d s^2_{10} ~({\rm type ~IIA}) = &&e^{5 \phi_{B0}/8}\, H'^{5/8} \,H^{1/4} \,
[ H^{-1} (- (d x^0)^2 +
(d x^1)^2) \nonumber\\
&&+ e^{-\phi_{B0}} H'^{-1} ( (d x^2)^2 + \cdots +
(d x^6)^2) + d y^i d y^i],
\label{eq:iiamsix}
\end{eqnarray}
with $i = 1, 2, 3$; the type IIA dilaton,
\begin{equation}
e^{\phi_A} = e^{ - \phi_{B0}/4} \,H'^{- 1/4}\, H^{- 1/2},
\label{eq:iiadsix}
\end{equation}
and the remaining non-vanishing fields,
\begin{eqnarray}
F_2 &=& n \,\frac{\sqrt{2} \kappa_0 Q_0^6}{\Omega_2}\, \epsilon_2,\nonumber\\
F_3 &=&- \Delta_{(m,n)}^{-1/2} \,e^{\phi_{B0}}\, (m - \chi_{B0} n)
\,d H^{-1}\wedge d x^0
\wedge d x^1, \nonumber\\
F_4' &=& - n \,\Delta_{(m,n)}^{- 1/2} \,e^{\phi_{B0}}\,
(m - \chi_{B0} n)\, H^{-1}\,
\frac{\sqrt{2} \kappa_0 Q_0^6}{\Omega_2}\, d x^0 \wedge d x^1
\wedge \epsilon_2,
\label{eq:iiarfsix}
\end{eqnarray}
Once again the Harmonic function $H'$ continues to be given by
Eq.\ (\ref{eq:nhf2}) but the Harmonic
function $H$ is
\begin{equation}
H = 1 + \frac{Q_6}{r},
\label{eq:iiahfsix}
\end{equation}
where $Q_6 = \Delta_{(m,n)}^{1/2} \sqrt{2} \kappa_0 Q_0^6 /\Omega_2$.
{\bf (F, D7) Bound State}: The various field configurations
in this case are described by the following Einstein metric,
\begin{eqnarray}
d s^2_{10}~ ({\rm type~ IIB}) =&& e^{3 \phi_{B0}/4}\, H'^{3/4}\, H^{1/4} \,
[ H^{-1} (- (d x^0)^2 +
(d x^1)^2)\nonumber\\
&& + e^{-\phi_{B0}} \, H'^{-1} ( (d x^2)^2 + (d x^3)^2 + \cdots +
(d x^7)^2) + d y^i d y^i],
\label{eq:iibmseven}
\end{eqnarray}
with $i = 1, 2$; the type IIB dilaton,
\begin{equation}
e^{\phi_B} = e^{- \phi_{B0}/2}\, H'^{ -1/2} \,H^{- 1/2},
\label{eq:iibdseven}
\end{equation}
and the remaining non-vanishing fields,
\begin{eqnarray}
d \chi &=& n \,\frac{\sqrt{2} \kappa_0 Q_0^7}{\Omega_1}\,
\epsilon_1,\nonumber\\
H_3 ~({\rm RR}) &=& - n \,\Delta_{(m,n)}^{- 1/2}\, e^{\phi_{B0}}\,
(m - \chi_{B0} n)\, H^{-1}\,
\frac{\sqrt{2} \kappa_0
Q_0^7}{\Omega_1} \,d x^0 \wedge d x^1 \wedge \epsilon_1, \nonumber\\
H_3 ~({\rm NSNS}) & = & - \Delta_{(m,n)}^{-1/2}\, e^{\phi_{B0}}\,
(m - \chi_{B0} n)\, d H^{-1}\wedge d x^0
\wedge d x^1.
\label{eq:iibrfseven}
\end{eqnarray}
Here the Harmonic function $H'$ is again given by Eq.\ (\ref{eq:nhf2})
but the Harmonic
function $H$ is now
\begin{equation}
H = 1 - Q_7 ~{\rm ln} r,
\label{eq:iibhfseven}
\end{equation}
where $Q_7 = \Delta_{(m,n)}^{1/2}\sqrt{2} \kappa_0 Q_0^7 /\Omega_1$.
A detail discussion of the meaning and the other properties of the
above solutions will be given in the following section. We, however,
continue this section to briefly indicate how to construct (W, Dp)
bound states in type II theories from the (F, Dp) bound states already
constructed.
So far, we have performed T-duality only along the transverse directions
of the F-strings to obtain (F, Dp) bound states from the known (F, D1) bound
state. But we note that following the prescription given in the previous
section, we can also T-dualize the above (F, Dp) solutions for $1\le p \le
7$ along $x^1$-direction of the F-strings.
This will give (W, Dp) bound state solutions for $0\le p \le 6$. We here
present
the Einstein metric, the dilaton and the Kaluza-Klein vector potential
for each of these
solutions. The remaining non-vanishing
fields can be obtained easily from the corresponding ones of
(F, Dp). For the (W, Dp) solutions we have the following Einstein metric,
\begin{eqnarray}
d s^2_{10} =&& e^{(p + 1) \phi_{B0}/8} \,H'^{( p + 1)/8}\,
[ - H^{ -1} (d x^0)^2
+ e^{-\phi_{B0}}\, H'^{-1} \, H (d x^1 + {\cal A}_0 d x^0)^2 \nonumber\\
&&+ e^{-\phi_{B0}}\, H'^{ -1} (
(d x^2)^2 + \cdots + (d x^p)^2) + d y^i dy^i],
\label{eq:gwdpm}
\end{eqnarray}
with $i = 1, 2, \cdots, (9 - p)$; the dilaton,
\begin{equation}
e^\phi = e^{(3 - p)\phi_{B0}/4} \,H'^{(3 - p)/4},
\label{eq:gwdpd}
\end{equation}
and the Kaluza-Klein vector potential
\begin{equation}
{\cal A}_0 = - \Delta_{(m,n)}^{- 1/2}\, e^{\phi_{B0}}
(m - \chi_{B0} n)\, H^{ - 1}.
\label{eq:gwdpv}
\end{equation}
In the above, the Harmonic function $H$ is given by,
\begin{equation}
H = \left\{\begin{array}{ll}
1 + \frac{Q_{p + 1}}{r^{6 - p}},& 0 \le p \le 5,\\
1 - Q_7~ \mbox{ln}~r, & p = 6, \end{array}
\right.
\end{equation}
and the Harmonic function $H'$ continues to be given by
Eq.\ (\ref{eq:nhf2}). All the other
quantities are already given before. Note that for $m = 0$
(also $\chi_{B0} = 0$ and for simplicity we
also set $\phi_{B0} = 0$\footnote{We will discuss this in the
following section.}),
we have $H' = H$
and ${\cal A}_0 = 0$. Then the above solution represents
an infinite periodic array of
Dp-branes along the $x^1$-axis. While for $n = 0$, $H' = 1$, and the
above solution
represents a gravitational wave propagating in $x^1$-direction with
isometries along
$(x^1, x^2, \cdots, x^p)$-directions. We will not discuss the (W, Dp)
bound states any
further in this paper and focus on (F, Dp) ones for the rest.
\section{Properties of (F, D$_p$) Bound States\protect\\}
\label{sec:pfdp}
The worldvolume picture for a (F, Dp) bound state described in
\cite{lurone} consists of
constant electric field lines flowing along, say, $x^1$-axis in the
Dp-brane
worldvolume. This picture applies for each of our spacetime configurations
of (F, Dp) given
in the previous section. In order to see that, one can examine that for
$m = 0$ (with
$\chi_{B0} = 0$ and $\phi_{B0} = 0$ for simplicity), the metric describes
Dp-branes lying along ($x^1, x^2, \cdots, x^p$)-directions
while for $n = 0$ (hence $H' = 1$), it describes F-strings lying along
the $x^1$-direction
with isometries along $(x^1,x^2, \cdots, x^p$)-directions. So,
this configuration indeed describes F-strings within the D-branes
in consistency with our
worldvolume picture. In order to make such an identification more precise,
we need to
calculate the charges carried by the Dp-brane and by the F-strings, and also
the mass per unit $p$-brane worldvolume, and compare them with what we
obtained in
\cite{lurone} based on the worldvolume approach.
Let us first calculate the RR charge for the Dp-brane in a (F, Dp)
bound state
configuration given in the previous section.
This can be done easily using either Eq.(3.9) or Eq.(3.10) depending
on whether the charge is electric-like or magnetic-like with the
corresponding explicit $(p + 2)$-form field strength
given in the previous section. We
have $e_p = n\, Q_0^p$ for an electric-like RR charge or
$g_p = n\, Q_0^p$ for a
magnetic-like RR charge. The integer $n$ originates from that of the
D-string in the
(F, D1) bound state. So the RR charge is automatically
quantized given the quantization of D-string charge, a well-known
fact that the
charge quantization for one extended object will imply charge
quantizations for the rest
of the extended objects in string/M theory. This
is consistent with our worldvolume result. Note that
$Q_0^p \,Q_0^{6 - p} = 2 \pi$.
As hinted earlier, Eq.\ (\ref{eq:ecd}) cannot be applied simply to calculate
the electric-like charge associated with the F-strings in (F, Dp).
The reason is, as
explained in \cite{lurone}, that the notation `F' in (F, Dp) means actually
an infinite
number of parallel strings in the bound state.
As mentioned in \cite{lurone}, there is one NS-string
(or equivalently $m$ F-strings) per $(2\pi)^{p -1}
\alpha'^{(p - 1)/2}$ of ($p - 1$)-dimensional area. We will show later in
this section that this is
indeed true.
In 1 + 3
dimensional electrostatics, we know that in order to use Gauss law to obtain
the charge per unit length for a uniform line distribution of charge,
we have to choose
a cylinder (with the line charge at the center) as the Gauss surface
rather than a 2-sphere as for a
point charge. For the present case, we therefore should choose the
integration in
Eq.\ (\ref{eq:ecd}) over $R^{p -1} \times S^{8 - p}$ rather than over
an asymptotic
$S^7$. We can therefore have charge per unit $(p -1)$-dimensional area
for the F-strings.
In order to give meaningful calculations for this quantity and for the
mass per unit Dp-brane worldvolume which will be used to determine the tension
for a (F, Dp) bound
state, we need a good asymptotic behavior for the metric such that the
$(p - 1)$-dimensional area and the worldvolume can be defined with
respect to certain frame
metric. This is also necessary for us to make comparisons with the results
obtained in \cite{lurone} based on the worldvolume analysis.
Let us take a close look at each
Einstein frame
metric for those (F, Dp) bound state configurations. They,
along with the corresponding
dilatons, can
actually be expressed in a unified way as
\begin{eqnarray}
d s^2 = &&e^{(p - 1) \phi_{B0}/8}\, H'^{(p - 1)/8}\, H^{1/4}\,
[ H^{ -1} (- (d x^0)^2 +
(d x^1)^2) \nonumber\\
&&+ e^{- \phi_{B0}}\, H'^{-1} \,( (d x^2)^2 + \cdots
+ (d x^p)^2) + d y^i d y^i],
\label{eq:gem}
\end{eqnarray}
for the metric with $i = 1, 2, \cdots, 9 - p$; and
\begin{equation}
e^\phi = e^{(5 - p)\phi_{B0}/4} \,H'^{(5 - p)/4} \, H^{ - 1/2}.
\label{eq:gd}
\end{equation}
for the dilaton. In the above, the Harmonic function $H$ is
\begin{equation}
H = 1 + \frac{Q_p}{r^{7 - p}},
\label{eq:ghf}
\end{equation}
with
\begin{equation}
Q_p = \frac{\Delta_{(m, n)}^{1/2} \sqrt{2} \kappa_0 Q_0^p}
{(7 - p) \Omega_{8 - p}},
\label{eq:gcc}
\end{equation}
and the Harmonic function $H'$ is given by Eq.\ (\ref{eq:nhf2}).
Note first that the constants
$\phi_{B0}$ and $\chi_{B0}$ appearing in all the above solutions
are no longer the
asymptotic values for the dilaton and the axion except for the original
(F, D1) bound state
configuration. Actually, the axion for $p = 3, 5, 7$ has nothing
to do with the constant
$\chi_{B0}$. In fact, $\chi_{B0}$ is the asymptotic value of the
$(p - 1)$-form RR gauge
potential
in the corresponding (F, Dp) configuration. The asymptotic value of
dilaton for every $p$
except for $p = 5$
is related to the original $\phi_{B0}$ but not equal.
This can be understood from the fact
that a T-duality transformation shifts the dilaton value\cite{bus}. From
the dilaton expression for each of the above solutions, we can see
that the asymptotic
value of the dilaton is reduced by a quarter of $\phi_{B0}$ for
every T-duality
transformation. For $p = 5$, the asymptotic value vanishes. Second,
even though
we start with an Einstein metric for (F, D1) which is
asymptotically Minkowski, none of
the metrics derived by T-duality transformations remains to be so.
Even worse, none of them
remains asymptotically as a constant scaling factor times Minskowski metric.
However,
there exists a constraint on the asymptotic behavior of the
Einstein metric for the
(F, D1) bound state configuration which preserves not only the
asymptotic behavior for each
of the derived Einstein metrics but also the asymptotic value for the
dilatons\footnote{This
remains true also for the (W, Dp) bound states.}. This turns
out to be the one which imposes the corresponding string metric
rather than the Einstein
metric to be asymptotically Minkowski. This constraint has been
used in the literature and
once again we see its significance. We will discuss possible
reasons behind this and other
related issues in a separate paper.
Once we make such a choice for the asymptotic metric,
Eq.\ (\ref{eq:gem}) and (\ref{eq:gd})
become
\begin{eqnarray}
d s^2 = && e^{- \phi_{B0}/2} \,H'^{(p - 1)/8} \,H^{1/4}\,\left[ H^{ -1}
(- (d x^0)^2 + (dx^1)^2)\right. \nonumber\\
&& \left. + H'^{-1} ( (d x^2)^2 + \cdots + (d x^p)^2) + d y^i d y^i
\right],
\label{eq:ngem}
\end{eqnarray}
and
\begin{equation}
e^\phi = e^{\phi_{B0}}\, H'^{(5 - p)/4}\, H^{ - 1/2}.
\label{eq:ngd}
\end{equation}
The Harmonic function $H$ is given by Eq.\ (\ref{eq:ghf}) but with
$Q_p \rightarrow e^{3 \phi_{B0}/2} Q_p$ with
$Q_p$ given in Eq.\ (\ref{eq:gcc}).
The expressions (4.5) and (4.6) have the above mentioned properties.
For example, the asymptotic value for the
dilaton is always $\phi_{B0}$, the one in the original (F, D1) bound state.
In what
follows, we simply use $\phi_0$ rather than $\phi_{B0}$ as the
asymptotic value for the
dilaton.
We give below the
explicit form of NSNS 3-from field strength\footnote{$G_3 = F_3$
in type IIA while
$G_3 = H_3 ~({\rm NSNS})$ in type IIB.} $G_3$ for the purpose of
calculating the
charge per unit $(p-1)$-dimensional area for F-strings which is
\begin{equation}
G_3 = - \Delta_{(m, n)}^{- 1/2} \,e^{\phi_{0}/2} \,(m - \chi_{B0} n) \,
d H^{ -1}\wedge d x^0
\wedge d x^1.
\label{eq:gthreef}
\end{equation}
As discussed above, the charge per unit $(p-1)$-dimensional area for the
F-strings should be calculated as,
\begin{equation}
e_1 = \frac{1}{\sqrt{2} \kappa_0} \int_{R^{p - 1}\times S^{8 - p}}
(e ^{ - \phi} \ast G_3
+ \cdots),
\label{eq:gcfs}
\end{equation}
where $\cdots$ indicates possible non-vanishing Chern-Simons terms.
Actually, the
Chern-Simons terms do contribute to this charge.
The charge $e_1$ itself must be infinite since we have an infinite
number of F-strings.
But we expect that $e_1 / A_{p -1}$ should be finite with
$A_{p - 1} = \int d x^2\wedge
\cdots \wedge d x^p$ the coordinate $(p -1)$-dimensional area. Indeed,
we find
\begin{equation}
|e_1|/A_{p -1} = m Q_0^p,
\label{eq:gcpfs}
\end{equation}
As mentioned in \cite{lurone}, we have one NSNS-string
(or $m$ F-strings) per
$(2\pi)^{p -1} \alpha'^{(p -1)/2}$ of ($p -1$)-dimensional area
from the worldvolume
point of view. So, in order to compare with our interpretation for the
F-strings along $x^1$-direction we must multiply the left side
of the above equation
with this $(p -1)$-dimensional area. Thus we obtain,
\begin{equation}
(2 \pi)^{p -1} \alpha'^{(p -1)/2} \frac{|e_1|}{ A_{p -1}}
= \sqrt{2} \kappa_0 \,m T_f,
\label{eq:conf}
\end{equation}
Now in order for the left hand side of Eq.(4.10) to represent the
charge of F-strings per $(2\pi)^{p-1} \alpha'^{(p-1)/2}$ of
$(p-1)$-dimensional area, we must replace the coordinate area $A_{p-1}$
by the one measured in string metric asymptotically. But, since we
have chosen our string metric to be asymptotically flat, they are the same.
Note also that according to our definition the charge $e_p$ of the
$p$-brane is related to the corresponding tension $T_p$ as $e_p =
{\sqrt 2} \kappa_0 T_p$ and therefore, Eq.(4.10) clearly states that
there are indeed $m$ F strings in the (F, Dp) bound states per $(2\pi)^{p-1}
\alpha'^{(p-1)/2}$ of $(p-1)$-dimensional area, confirming our interpretation.
There is another way to obtain the above result which gives a direct link
between the
spacetime F-strings and the worldvolume electric flux lines.
This will be important for
our discussion of the $AdS_5/CFT_4$ correspondence given in section 5.
The low-energy
effective field theory on a Dp brane worldvolume contains a coupling,
\begin{equation}
\frac{1}{T^p_0 g} \int d^{p + 1} \sigma B_{\mu\nu} (2\pi \alpha' F^{\mu\nu}),
\label{eq:coupling}
\end{equation}
where $B_2$ is the pull-back of the NSNS 2-form potential on the worldvolume,
$F$ is the worldvolume gauge field strength\footnote{Strictly
speaking, we should use ${\cal F} = F - B_2$ instead.},
$T_0^p = 1/[(2\pi)^p \alpha'^{(p + 1)/2}]$ is the $p$-brane tension units
and $g = e^{\phi_0}$ is
the string coupling. Because of this term the equation of motion
takes the form (in terms of our
conventions)\footnote{Here,
we do not include an F-string source to the equation
of motion since we are considering an infinite number
of parallel F-strings in the bulk.}
\begin{equation}
d (e^{- \phi}\ast G_3 + \cdots ) = \frac{ 2 \,\kappa_0^2}
{ g \,T_f \,T^p_0} \ast F \wedge \delta^{9 - p},
\label{eq:ceom}
\end{equation}
where $\ast$ in $\ast G_3$ denotes the Hodge dual in $D = 10$ spacetime while
the $\ast$ in $\ast F$ denotes the Hodge dual within the
$(p + 1)$-dimensional worldvolume.
In the above, $\delta^{9 - p}$ denotes a $(9 - p)$-form delta function
on the Dp brane
worldvolume. We use this equation for the purpose of relating
the F-string charge to the
electric flux of gauge field on the worldvolume.
(Note that a similar equation has been used in \cite{str} for the
charge conservation.) In order to get the charge of F-strings in (F, Dp)
bound state, we should
integrate the above equation on $R^{p - 1} \times R^{9 - p}$. We have
\begin{eqnarray}
\int_{R^{p - 1} \times R^{9 - p}} d (e^{- \phi} \ast G_3 + \cdots) &=&
\int_{R^{p - 1} \times S^{8 - p}} (e^{-\phi} \ast G_3 + \cdots ),\nonumber\\
&=& \frac{ 2\, \kappa_0^2}{ g \,T_f \,T^p_0} \int_{R^{ p - 1}} \ast F.
\label{eq: iceom}
\end{eqnarray}
This gives,
\begin{eqnarray}
e_1 &=& \frac{1}{\sqrt{2} \kappa_0} \int_{R^{p - 1}\times S^{8 - p}}
(e ^{ - \phi} \ast G_3 + \cdots), \nonumber\\
&=& \frac{ \sqrt{2} \kappa_0}{ g T_f T^p_0} F_{01} A_{p - 1},
\label{eq:cfr}
\end{eqnarray}
where we have taken $F_{01}$ as the only non-vanishing constant component
of the $U(1)$ gauge field strength on the worldvolume of Dp-brane.
If we denote $Q_1$ as the
charge of F-strings per $(2\pi)^{p -1} \alpha'^{(p -1)/2}$ area, we
have
\begin{equation}
\frac{Q_1}{\sqrt{2}\, \kappa_0} = F_{01} / g,
\label{eq:cfrone}
\end{equation}
where $Q_1/\sqrt{2} \kappa_0$ is the total tension of these F-strings.
So we establish the precise
relationship between the F-strings and the electric flux lines. If we take
$Q_1/\sqrt{2} \kappa_0 = m T_f$ from the above calculated value, we
have $F_{01} = g m T_f$ which
agrees precisely with what we obtained in \cite{lurone}. Certainly, the other
way around is also true.
We now use the metric Eq.\ (\ref{eq:ngem}) to calculate the mass per unit
$p$-brane volume
for a (F, Dp) bound state and compare the result with what we obtained
in \cite{lurone}.
In doing so, we need to generalize the ADM formula
in [18] to accommodate the following
metric,
\begin{equation}
d s^2 = - A(r) dt^2 + B(r) d r^2 + r^2 C (r) d \Omega_{D - p - 2} +
D (r) \delta_{ij} d x^i d x^j + E (r) \delta_{kl} d x^k d x^l,
\label{eq:gm}
\end{equation}
where $D$ is the spacetime dimensions, $p$ is the spatial dimensions of a
$p$-brane, and
indices $(i, j = 1, 2, \cdots, m)$ and $(k, l = m + 1, \cdots, p)$. If
$ A(\infty) = B(\infty) = C(\infty) = D (\infty) = E (\infty) = a_p$
(with $a_p$ a constant)
[19], we have the ADM mass per unit $p$-brane volume as
\begin{eqnarray}
M_p = - \frac{\Omega_{\tilde{d} + 1}}{2 \kappa^2_0} [ &&
(\tilde{d} + 1) \,r^{\tilde{d} + 1}\partial_r C(r) + m \,r^{\tilde{d}
+ 1} \partial_r D (r)
+ (p - m) \,r^{\tilde{d} + 1} \partial_r E (r) \nonumber\\
&&- (\tilde{d} + 1)\, r^{\tilde{d}}
\left(B (r) - C (r)\right) ]_{r \rightarrow \infty},
\label{eq:gadm}
\end{eqnarray}
with $\tilde{d} = D - p - 3$.
Applying the above formula to our case, we get the mass per unit
$p$-brane volume for
a (F, Dp) bound state as
\begin{equation}
M_p = g \sqrt{n^2 + g^2 (m - \chi_{B0} n)^2} ~T^0_p,
\label{eq:gmass}
\end{equation}
where we have used
$Q_p = \Delta_{(m,n)}^{1/2} e^{3\phi_{0}/2} \sqrt{2}\kappa_0 Q_0^p/[(7 - p)
\Omega_{8 - p}]$,
the expressions for $Q_0^p$ from Eq.\ (\ref{eq:puc}) and
the $\Delta$-factor from
Eq.\ (\ref{eq:df}), respectively. Here $T^p_0 = 1/[(2\pi)^p
\alpha'^{(p + 1)/2}]$ is the $p$-brane
tension units and $g = e^{\phi_{0}}$ is the string coupling. We can relate the
$M_p$ and the tension $T_p (m, n)$ of (F, Dp) in many ways.
{}For example, they can be related by looking at the scaling behavior of the
energy-momentum
tensor due to the contribution of the Dp-brane worldvolume action
under a constant rescaling
of the spacetime metric. But we here use a very simple approach.
We know that $M_p$ must be
proportional to $T_p (m, n)$ and the proportionality constant must be
some power of the constant
$a_p$, i.e., $M_p = a_p^\alpha T_p (m, n)$ with $\alpha$
an as yet undetermined constant.
Note that here $a_p = g^{- 1/2}$. When $m = 0, \chi_{B0} = 0$,
we know $T_p (0, n) = n T^p_0/ g$ and from Eq.\ (\ref{eq:gadm})
we also have
$M_p = g n T^p_0$.
So we
have $\alpha = - 4$ and we get $M_p = g^2 T_p (m, n) $ in general.
Using this relation and Eq.(4.14) we find
the tension $T_p (m, n)$ for the (F, Dp) bound state as
\begin{equation}
T_p (m, n) = \frac{1}{g} \sqrt{n^2 + (m - \chi_{B0} n)^2 g^2}~ T^p_0,
\label{eq:gtension}
\end{equation}
which agrees precisely with what we obtained in \cite{lurone} for $\chi_{B0} = 0$. This also
shows that (F, Dp) is a non-threshold bound state.
\section{Decoupling Limit for the (F, D3) bound state\protect\\}
\label{sec:ads/cft}
The properties of the (F, Dp) bound states studied in the
previous section are either the
BPS ones or the consequences of
BPS properties. Therefore, it is not surprising that we find agreements
in both the spacetime and
the worldvolume approaches even though, strictly speaking,
a flat background is
always assumed in the worldvolume study for these bound states
in \cite{lurone}.
These calculations clearly demonstrate that our spacetime
bound state configurations (F, Dp) are identical to those bound states
obtained from the
worldvolume study in\cite{lurone}. The F-strings in (F, Dp) should be
identified with the
electric flux or field lines in the corresponding gauge theory
living on the worldvolume.
Now the question is: Can the same $AdS/CFT$ correspondences
conjectured by Maldacena
in \cite{mal} for $p = 3$ and by Itzhaki et al in \cite{itzmsy}
for $p \neq 3$
be proposed in a
similar fashion based on the (F, Dp) bound states rather than
on the simple Dp branes? If so,
can we learn anything new? We will examine these for the $p = 3$ case
in the following.
The analysis for $p \neq 3$
can be made similarly as for the $p = 3$ case along the
line given in \cite{itzmsy}.
In general, when the so-called D-branes appear in string theory,
we need to consider not only
the modes that propagate in the bulk but also the modes that
propagate on the D-branes. The modes
propagating on the D-branes are associated with open strings ending
on the D-branes. In general,
these modes interact not only among themselves but also with the
modes propagating in the bulk.
However, there exists a limit that decouples the
modes propagating on the branes from the modes propagating in the
bulk and is also typically
a low energy limit. The latter says that the modes on the brane are
just the massless ones of
the open strings. In this limit, the D-brane theory becomes the
corresponding SYM theory (for
$p \le 3$).
In the case of $p = 3$, the D3 brane theory is described by
${\cal N} = 4$ SYM
in $1 + 3$ dimensions under this limit. But the D3 brane configuration is
also described in terms of the metric
and other fields in the bulk. The so-called decoupling or field
theory limit is
\begin{equation}
g_{\rm YM}^2 = 2\pi g = {\rm fixed},~\qquad \alpha' \rightarrow 0,
\label{eq:dcl}
\end{equation}
where $g = e^{\phi_0}$ is the string coupling constant and
$g_{\rm YM}$ is the Yang-Mills coupling constant. Note that
the ten dimensional Newton constant $\kappa^2 \sim g^2 \alpha'^4$
vanishes in this limit as it should be.
Under this limit, the BPS D3-brane configuration in the bulk becomes
flat Minkowski for any fixed non-zero
$r$. Therefore, to have something
nontrivial, we need to consider an additional limit,
\begin{equation}
U \equiv \frac{r}{\alpha'} = {\rm fixed},~\qquad \alpha' \rightarrow 0.
\label{eq:esl}
\end{equation}
The above says that we keep the mass of the stretched strings
between D3 branes fixed. This $U$ actually
sets the energy scale in the field theory since it is the expectation
value of the Higgs. This also
implies that we are considering finite energy configurations
in the field theory. As discussed in
\cite{mal}, if we consider $n$ parallel D3 branes, then the above
decoupling limits say that we
are bringing the branes together but the Higgs expectation value
corresponding to this separation
remains fixed. The resulting theory on the brane is ${\cal N} = 4$ $U(n)$
SYM theory in $1 + 3$
dimensions. In what follows, we simply call
Eqs.\ (\ref{eq:dcl}) and \ (\ref{eq:esl}) as the decoupling
limit.
Under the decoupling limit, the string metric describing
the D3 branes becomes
$AdS_5 \times S^5$. The isometry of $AdS_5$ is $SO(2, 4)$
and this is also the conformal group
in $1 + 3$ dimensions. We also have isometries of $S^5$ as
$SO(6)\sim SU(4)$. This symmetry is
identical to the R-symmetry of the ${\cal N} = 4$ SYM. After
including fermionic generators
required by supersymmetry, we have the full isometry supergroup
$SU(2, 2\mid 4)$ for the
$AdS_5 \times S^5$ background, which is identical to the
${\cal N} = 4$ superconformal group.
We would like to emphasize that we have the conformal symmetry
$SO(2,4)$ only at $U = 0$. For any fixed
non-zero $U$, this symmetry is spontaneously broken since
$U$ represents the expectation value of the
Higgs.
To validate the configuration of D3 branes as a stringy one,
the large $n$ limit has to be taken given
the fixed but small string coupling\footnote{For large $g$,
the D-string frame is chosen.}.
It is well-known that there is an SL(2,Z) strong-weak duality symmetry
in both type IIB supergravity and
${\cal N} = 4$ SYM. These symmetry identifications, among other things,
led Maldacena to conjecture
that the large $n$, ${\cal N } = 4$ SYM theory with gauge group
$U(n)$ is actually equivalent to the
ten dimensional type IIB supergravity on $AdS_5 \times S^5$.
However, the supergravity itself is not
a consistent quantum theory. On the other hand, the SYM is a
unitary quantum theory. Moreover, the above
symmetry identifications are independent of the large $n$ limit
(even though it is needed to validate the
solution). These two facts
led Maldacena further to conjecture that type IIB string on
$AdS_5 \times S^5$ is equivalent to
the ${\cal N} = 4$ SYM theory for a general $n$.
Let us now apply the same process to the (F, D3) configuration
obtained in section 3, i.e.,
taking the decoupling limit given by Eqs.\ (\ref{eq:dcl}) and \ (\ref{eq:esl}). In other words,
we should
examine the near-horizon geometry of (F, D3) configuration with the
string coupling $g$ fixed\footnote{For simplicity, we set $\chi_{B0} = 0$
from now on.}.
The
near horizon string-frame metric in this case is
\begin{eqnarray}
d s^2 =&& \alpha' \left[ \frac{U^2}{\sqrt{4\pi n g}} \left(\frac{ n^2
e^{-\phi_0}}{\Delta_{(m,n)}}\right)^{3/4}
( - (d x^0)^2 + (d x^1)^2 ) \right.\nonumber\\
&& + \frac{U^2}{\sqrt{4\pi n g}} \left(
\frac{ n^2 e^{-\phi_0}}{\Delta_{(m,n)}}\right)^{ - 1/4}
((d x^2)^2 + (d x^3)^2)
\nonumber\\
&& \left. + \sqrt{4 \pi n g} \left(\frac{ n^2
e^{-\phi_0}}{\Delta_{(m,n)}}\right)^{1/4} \left(\frac{d U^2}{U^2} +
d \Omega_5^2 \right) \right],
\label{eq:nhg}
\end{eqnarray}
and the dilaton is
\begin{equation}
e^\phi = \left(n^2 e^{-\phi_0} /\Delta_{(m, n)}\right)^{1/2} e^{\phi_0}.
\label{eq:dthree}
\end{equation}
It is clear from (5.3) and (5.4) that unlike the simple D3 brane, the
near-horizon geometry in this case is
not automatically $AdS_5 \times S^5$ and the effective string coupling,
$e^\phi$,
is still a finite constant but is less than the string
coupling\footnote{The dilaton
for (F, D3) is actually not a constant in general, nevertheless
it is bounded as
$ \left(n^2 e^{-\phi_0} /\Delta_{(m, n)}\right)^{1/2}
e^{\phi_0}< e^\phi < e^{\phi_0}$,
where the upper bound is obtained at $r = \infty$ and the
lower bound at $r = 0$.} $g$.
The above features may be expected given the appearance of an infinite
number of F-strings in
(F, D3) and the non-threshold bound state nature of this configuration.
On the other hand,
in the decoupling
limit, we expect that D3 brane can be described equivalently
either by the SYM theory
on the brane or by the type IIB supergravity (or string)
in the bulk. Depending on whether we consider
excitations with respect to the simple D3-brane vacuum or the
(F, D3) bound state (chosen as a new vacuum),
we have two descriptions which are different in appearance but
probably equivalent in essence. Let us first
discuss the possible underlying IIB string/SYM correspondence
for each of the above descriptions in
the strong sense, i.e., not taking the large $n$ limit. We will discuss
the large $n$ limit in the end.
If we choose the simple BPS D3 brane configuration in the bulk
or equivalently no excitations in the SYM
on the brane as the vacuum, then the F-strings in the bulk or
equivalently the electric flux lines
in the
SYM should be formed due to excitations with respect to this vacuum.
We know that
the energy per unit 3-brane volume associated with the F-strings
in the bulk or the energy density
associated with
the electric flux lines on the brane is finite. This implies that we
have an infinite amount of energy
associated with either the F-strings or the electric flux lines.
In other words, we should have $U = \infty$,
which implies that the original conformal symmetry group
$SO(2,4)$ must be broken spontaneously to some
smaller symmetry group consisting of at most some translational
and rotational symmetries. It is not difficult
to figure out the origin of $U = \infty$ if we recall the method we used
in \cite{lurone}
to find the (F, D3) bound state. As discussed in \cite{lurone},
the charge conservation must imply that the
infinite parallel F-strings in (F, D3) originate from F-strings
along the radial coordinate $r$ ending on
the $x^2x^3$-plane placed at $x^1 = - \infty$
\footnote{This is chosen in order to preserve the SO(6)
isometry of $S^5$.}.
These F-strings ending on $x^2x^3$ plane are the open strings,
each semi-infinitely long, giving $U = \infty$. Inside the bulk, we
have F-strings along $x^1$-axis, therefore we expect an $SO(1,1)$
symmetry. Because of the ending or the charges on
$x^2x^3$-plane placed at $x^1 = - \infty$, we expect that
$SO(2,4) \rightarrow SO(1, 1) \times SO(2 )$ for this
near-horizon geometry. Thus we find that we actually have $SO(1,1)
\times SO(2 ) \times SO(6)$ isometries for
the near-horizon geometry of the (F, D3) bound state for the
choice of vacuum. On the SYM side, we have
the same answer. Dilation and the special conformal transformations
are broken spontaneously for the same reason
mentioned above. Only the subgroup $SO(1,1) \times SO(2)$ of
$SO(2, 4)$ will leave $F_{01} = g m T_f$ invariant.
So, following the similar argument as for the $AdS_5/CFT_4$ correspondence,
we may suggest that the type IIB string on this
geometry is equivalent to the ${\cal N } = 4$ SYM theory in a state
with quantized electric flux lines.
Let us
investigate how the (F, D3) bound state emerges as our vacuum
for the second description.
Examining carefully the metric and the dilaton, we can see that
$ n^2 e^{-\phi_0} /\Delta_{(m, n)}$
factor is the source which causes
the near-horizon geometry to be not automatically $AdS_5 \times S^5$
and the effective string coupling
to be less than the string coupling $g$. Explicitly, we have this factor
\begin{equation}
\frac{n^2 e^{-\phi_0}}{ \Delta_{(m,n)}} =
\left(1 + \left(\frac{g m}{n}\right)^2
\right)^{-1}.
\label{eq:fac}
\end{equation}
Now if we rescale the
coordinate time $x^0$ and the coordinate $x^1$ by a factor
$ \left(n^2 e^{-\phi_0} /\Delta_{(m, n)}\right)^{-1/2}$,
then in terms of the new coordinates the
near-horizon geometry
is $AdS_5 \times S^5$. So, if we rewrite the
metric in terms of the effective string coupling
\begin{equation}
g_{\rm eff} \equiv e^\phi = \frac{g}{\left(1 +
\left(\frac{g m}{n}\right)^2
\right)^{1/2}},
\label{eq:esc}
\end{equation}
we have
\begin{equation}
d s^2 = \alpha' \left[ \frac{U^2}{\sqrt{4\pi n g_{\rm eff}}}
\left( - (d x^0)^2 + (d x^1)^2 + (d x^2)^2 + (d x^3)^2 \right) +
\sqrt{4 \pi n g_{\rm eff}} \left(\frac{d U^2}{U^2} +
d \Omega_5^2 \right) \right].
\label{eq:enhg}
\end{equation}
The radius for the $AdS_5$ and the $S^5$ is now given as
\begin{equation}
R^2/\alpha' = \sqrt{4\pi n g_{\rm eff}},
\label{eq:radius}
\end{equation}
The information about the F-strings in the (F, D3)
bound state disappears in the above metric. Since the resulting
near-horizon geometry differs from
that of a simple D3 brane with 5-form flux $n$ only in the string
coupling, we expect that the effects of
the F-strings are encoded in the effective string coupling $g_{\rm eff}$.
We will see that this is indeed true.
With respect to the new coordinates, we have $SO(2,4) \times SO(6)$
isometries in the bulk
at least in appearance. How can we reconcile this with what we just
discussed above on the SYM
side? Moreover, what is the physics behind such a rescaling?
In the case of simple D3 branes, the string
coupling $g = e^{\phi_0}$ remains the same whether we are at
$r = \infty$ or at $r = 0$.
This fact enables us to fix the gauge coupling
as $g_{\rm YM}^2 = 2 \pi g$ for the SYM theory describing
the D3 branes in the decoupling limit.
We do not expect that such a relation will be modified for
any finite energy excitations\footnote{
But it could be modified
if the energy involved is infinite, for example, the case we are presently
studying. If this happens, the SYM theory
may no longer be valid in describing the underlying physics in general and
string theory should be used instead.
However, our discussion above for the first description may still
be valid since it is based on the BPS
(F, D3) configuration.}. However, the story here is different.
The gauge coupling should, in the present case, be
related to $g_{\rm eff}$ rather than
$g$ as $g_{\rm YM}^2 = 2\pi g_{\rm eff}$.
With respect to the metric in Eq.\ (\ref{eq:enhg}), we may say that
a simple D3 brane with 5-form flux $n$ plays
as an ``effective" vacuum configuration in the bulk with the
string coupling $g_{\rm eff}$. This simple D3 brane
is also described by a ${\cal N} = 4$ $U(n)$ SYM theory in 1 + 3
dimensions with gauge coupling
$g_{\rm YM}^2 = 2\pi g_{\rm eff}$. This is just the usual $AdS_5/CFT_4$
correspondence except we have a new
coupling constant.
Let us dig out the mystery behind the above ``effective" vacuum.
We expect that the tension for the simple
D3 brane is $n T^3_0 / g_{\rm eff}$. If we use the explicit
expression for $g_{\rm eff}$ from
Eq.\ (\ref{eq:esc}), this tension is
\begin{equation}
\frac{n T_0^3} {g_{\rm eff}} = \frac{T_0^3}{g} \sqrt{n^2 + g^2 m^2},
\label{eq:ident}
\end{equation}
which is just the tension for the (F, D3) bound state. So this
``effective" vacuum is nothing but
the (F, D3) bound state. As anticipated, the effects of the F-strings
in (F, D3)
is encoded into the string coupling constant.
On the SYM side, we have already hinted that the gauge coupling
is now given by
$g_{\rm YM}^2 = 2 \pi g_{\rm eff}$. In the linear approximation,
\begin{equation}
\frac{1} {g_{\rm eff}} = \frac{1}{g} + \frac{1}{2 g}
\left(\frac{ g m}{n}\right)^2,
\label{eq:lgc}
\end{equation}
where the second term for $n = 1$ is nothing but the contribution
from the energy density of the electric
flux lines. In other words, the effect of the energy of the
quantized electric flux lines is absorbed into the
gauge coupling. This is in accordance with our spacetime picture
just discussed.
Therefore, the resulting SYM theory should
be the one describing the excitations with respect to the (F, D3)
bound state. The superconformal symmetry
should then be
restored. We therefore have a consistent picture both in the
bulk and on the brane in the decoupling limit.
Our study lends further support for Maldacena's $AdS_5/CFT_4$ correspondence
and shows that it holds true
even for the
non-trivial D3 brane configurations. The new input for this correspondence
is that we should use the effective string coupling constant
instead of the usual one.
The two descriptions studied may be equivalent but the picture
for the first one is surely complicated. The
reason for this is our improper choice of the vacuum state.
If we take the large $n$ limit, i.e.,
$n \gg m$ for fixed $g$, the effective string coupling becomes the
usual string coupling $g$. The F-strings in the
(F, D3) now play a minor role with respect to the D3 branes which
can be seen from the tension formula
$T_3 (m, n) = n/g \left[ 1 + (m /n)^2 g^2 /2\right] T^3_0$.
On the SYM side, on the other hand, we have the gauge coupling
$g_{\rm YM}^2 = 2 \pi g (1 - (m/n)^2 g^2 /2)$ which implies that
electric flux lines plays the minor role.
Thus the $AdS_5/CFT_4$
correspondence with respect to (F, D3) bound state is simply reduced
to that of the
simple D3 brane configuration. We also expect that the full
isometry supergroup $SU(2,2\mid 4)$ in the bulk
and the superconformal symmetry in the SYM theory are restored under
this limit in our first description above.
This description is therefore also reduced to the usual $AdS_5/CFT_4$
correspondence. This large $n$ limit also
validates the (F, D3) configuration as a stringy one.
Let us now comment on the effective string coupling $g_{\rm eff}$ given by
Eq.\ (\ref{eq:esc}) in the near-horizon region. It is clearly quantized
in terms of relatively
prime integers $n$ and $m$ and is always less than the string
coupling $g$. In particular, in the limit $m \gg n$,
it becomes $g_{\rm eff} = n /m
\ll 1$, independent of the string coupling $g$. In other words, the
effective string coupling in the
near-horizon region is completely determined by integer $n$, the
3-brane charge, and $m$,
the number of F-strings per $(2\pi)^2 \alpha'$ area over the
$x^2x^3$-plane, in this limit.
We will continue this study in a more general
D3-brane configuration, namely, ((F, D1), D3) bound state in a
forthcoming paper \cite{lurthree}.
\section{Conclusion\protect\\}
\label{sec:c}
To summarize, we have constructed, in this paper, explicitly the (F, Dp)
non-threshold bound state configurations for
$2 \le p \le 7$, starting from the known (F, D1) configuration of Schwarz
by T-duality
transformation along the transverse directions of the strings. We have
also presented the solutions
for (W, Dp) non-threshold bound state configurations for
$0 \le p \le 6$ by T-duality
transformations on the newly obtained (F, Dp) ones along the
longitudinal direction of F-strings
in (F, Dp). We have shown explicitly that there are $m$ F-strings per
$(2\pi)^{p - 1} \alpha'^{(p - 1)/2}$ of $(p - 1)$-dimensional area
which agrees with our previous result
based on the worldvolume study. We have also calculated the tensions
for the (F, Dp) bound states which
once again agree with our previous results. These two facts
confirm our assertion made in our earlier paper that the bound state
studied in \cite{lurone}
can indeed be identified with the (F, Dp) bound states here. All of
these bound states preserve one
half of the spacetime supersymmetries.
From the D-brane worldvolume point of view, each of these
bound states consists of
a Dp-brane carrying certain units of quantized electric flux or
field line. But from the
spacetime point of view, we have type II strings (either type IIA or
IIB strings) lying along
one direction in the corresponding Dp-brane worldvolume. In this paper,
we have shown that for
each of the bound states considered, the
type II strings can be identified with the electric flux lines
in the gauge theory
living on the corresponding Dp brane worldvolume. In the $p = 3$ case,
we have studied the corresponding
decoupling limit and found that the $AdS_5/CFT_4$ correspondence may
still hold true but now with
respect to the (F, D3) bound state with an effective string coupling constant.
The F-strings in the (F, D3) bound
state or the electric flux lines in the corresponding SYM theory play the
role of reducing the respective
coupling constant in general. The string coupling in the near-horizon
region is found to be quantized in terms of the relatively
prime integers $n$, the 3-brane charge, and $m$, the number of
F-strings per $(2 \pi)^2 \alpha'$ area over the
$x^2x^3$-plane. It becomes independent of the asymptotic
string coupling $g$ in the limit $m \gg n$.
\acknowledgments
We would like to thank Mike Duff for reading the manuscript and Chris Pope
for discussions. JXL acknowledges the support of NSF Grant PHY-9722090.
|
\section{Introduction}
\label{intro}
Soft gamma--ray repeaters (SGRs) are a class of astrophysical
sources that emit bursts of high energy x--ray and gamma--ray
radiation which are among the most energetic events in the Galaxy.
The apparent association of their positions with supernova remnants
and the detection of pulse periods in their nonbursting emission
strongly suggest that the SGRs are young neutron stars (e.g. Mazets
et al. 1979, and review by Rothschild 1995). The SGRs may also be
related to the anomalous X--ray pulsars (AXPs: \cite{mer97}), which
have comparable long ($>$ few second) periods. The observed SGR burst
energies, assuming isotropic emission, range from typical values of
$\sim 10^{41}$ ergs to as much as $10^{44}$ ergs in rare giant flares,
such as that of 5 March 1979 from the SGR 0529--66 in the Large
Magellanic Cloud. Suggested energy sources for these bursts have
included, i) the rotational energy of the neutron star, $\sim 10^{45}
{(P/3.1\,\rm{s})}^{-2}$ ergs, where $P$ is the spin period, which might
be tapped by pulsar glitches (e.g. \cite{baym71}), ii) the magnetic
field energy $\sim 10^{44}{(B/B_{q})}^{2}$ ergs of {\it magnetars} with
surface magnetic fields much greater than the quantum critical field
$B_{q}=m_{e}^{2}c^{3}/ e\hbar\approx 4.4\times 10^{13}$ G tapped by
magnetic-stress driven crustal quakes and magnetic reconnection
(\cite{thompson95}), and iii) the gravitational binding energy of
the neutron star, $\sim 10^{53}$ ergs, tapped by quakes (e.g Ramaty
et al. 1980), and driven by plate tectonics (\cite{ruderman91}).
Recent measurements of the rapid spindown rates of the SGR pulsars
have been taken (e.g. \cite{kouv98}, 1999) as evidence for the
magnetar hypothesis, in which the magnetic energy of the neutron
star exceeds the rotational energy. Pulsations have been observed
from three of the SGRs: SGR 0526--66 ($8$ s: \cite{mazets79}),
SGR 1806--20 ($7.47$ s: \cite{kouv98}), and SGR 1900+14 ($5.16$
s: \cite{hurley99b}). The period derivatives ($\dot{P}$) of these
pulsars have been found by either direct measurement (SGRs 1806--20
and 1900+14) or by $\dot{P}=0.5 P/t_{snr}$, where $P$ is the pulse
period and $t_{snr}$ is the estimated age of the associated supernova
remnant (SGR 0526--66). If the spindown is driven by magnetic dipole
radiation from an orthogonally rotating vacuum magnetic dipole, it
can be shown (\cite{pacini68}) that the surface magnetic field is
given by $B_{0}\approx 3.2\times 10^{19}\sqrt{P\dot{P}}$ G, which
would yield surface magnetic fields of $6\times 10^{14}$, $8\times
10^{14}$, and $5\times 10^{14}$ G for SGRs 0526--66
(\cite{thompson95}), 1806--20 (\cite{kouv98}), and 1900+14
(\cite{kouv99}), respectively. Here we present {\it RXTE}
observations, however, which suggest that the spindown rate
of SGR 1900+14 is due to torques other than those provided by
the magnetic field, and thus does not provide evidence of a
supercritical surface dipole field.
\section{Observations \& Analysis}
\label{obervations}
SGR 1900+14 was observed by the Proportional Counter Array (PCA)
and High Energy X--ray Timing Experiment (HEXTE) instruments aboard
the {\it Rossi X--ray Timing Explorer} on a number of occasions
during the period September 4-18, 1996. The total exposure time was
$\sim 47$ ks, with a temporal baseline of $15.4$ days. For the
first $22$ ks, {\it RXTE} was pointed at a position RA (J2000)$=
286^{\circ}.82$ and Dec (J2000)$=9^{\circ}.32$, which is $\sim 48
\arcsec$ from the precise VLA position of SGR 1900+14 (\cite{frail99}),
but well inside the $1^{\circ}$ FWHM field of view of the {\it RXTE}
pointed instruments. Midway through the observations, the pointing
position was changed to exclude the bright $438$ s binary x--ray
pulsar 4U 1907+09 (\cite{zand98}) from the field of view. The
second half of the observation ($25$ ks) was then conducted at
the pointing position RA$=286^{\circ}.43$ and Dec$=8^{\circ}.98$,
which is $\sim 0^{\circ}.35$ from the position of the SGR. As luck
would have it, this field also contained a relatively bright
confusing source, the $89$ s transient x--ray pulsar XTE J1906+09,
which was discovered during the observation ({\cite{marsden98}).
Finally, the Galactic Ridge emission is also a significant contributor
to the x--ray flux in the {\it RXTE} field of view (\cite{valinia98}),
due to the low Galactic latitude of SGR 1900+14 ($b\sim 0^{\circ}.75$).
Because of these complications, we do not attempt to determine the
x--ray spectrum of the SGR with the {\it RXTE} data, and instead
concentrate on the temporal analysis. For information on the x--ray
spectrum of the source, the reader is referred to Hurley et al.
(1999b), Kouveliotou et al. (1999), and Murakami et al. (1999).
The pointed x--ray instruments aboard {\it RXTE} are the High Energy
X--ray Timing Experiment (HEXTE) and the Proportional Counter Array
(PCA). HEXTE consists of two clusters of collimated NaI/CsI phoswich
detectors with a total net area of $\sim 1600$ cm$^{2}$ and and
effective energy range of $\sim 15-250$ keV (\cite{rothschild98}).
The PCA instrument consists of five collimated Xenon proportional
counter detectors with a total net area of $7000$ cm$^{2}$ and an
effective energy range of $2-60$ keV (\cite{jahoda96}). The uncertainty
in the timing of x--ray photons by the PCA and HEXTE is $<<1$ ms (\cite{rots98}), and is therefore negligible in the temporal analysis
presented here.
The PCA and HEXTE photon times were corrected to the Solar System
barycenter using the JPL DE200 ephemeris and the SGR coordinates RA(J2000)$=19^{\rm{h}} 07^{\rm{m}} 14\fs 33$ and Dec(J2000)$=+09^{\circ}19\arcmin 20\arcsec.1$ (\cite{frail99}).
The PCA data were searched for pulsations using the chi-squared
folding method, which calculates the value of chi-squared for a
pulsar lightcurve (versus a constant rate) folded on a range of
trial pulsar periods. Here the pulse phase $\phi$ for a given
photon time $t$ is defined by the relation $\phi(t)=f(t-t_{0})+
{1\over 2}\dot{f}{(t-t_{0})}^{2}$, where the pulsar frequency $f$
and frequency derivative $\dot{f}$ are related to the period $P$
and period derivative $\dot{P}$ by the expressions $P=1/f$ and
$\dot{P}=-\dot{f}P^{2}$. A maximum value of chi-squared occurs
when the data are folded on the true pulsar period and period
derivative.
The PCA data were initially searched for pulsations using a range
of $\sim 500$ periods about $5.153642$ s, the SGR 1900+14 period
predicted from the timing ephemeris given in Kouveliotou et al.
(1999). A significant chi-squared peak was seen, and a finer search
was then conducted on a grid in $P-\dot{P}$ space around the peak, for
a broad range of $\dot{P}$ including the value of $\dot{P}\sim 10^{-10}$
s s$^{-1}$ found by Kouveliotou et al. (1999). The results of the grid
search are shown in Figure $1$. To estimate the confidence regions of
$P$ and $\dot{P}$ indicated by the peak in chi-squared, we folded the
$2-10$ keV PCA data with $P$ ($\dot{P}$) values slightly displaced from
the peak value, while holding $\dot{P}$ ($P$) fixed at its peak value.
The resultant lightcurves were then compared to a template lightcurve
using the chi-squared test, and the $90\%$ confidence contours were
calculated using the chi-squared probability distribution. A folding
time midway through the {\it RXTE} observation was used throughout
the analysis to minimize correlations between $P$ and $\dot{P}$.
Using this analysis, we obtain a timing solution of $P=5.1558199\pm
0.0000029$ s and $\dot{P}=(6.0\pm 1.0)\times 10^{-11}$ s s$^{-1}$,
referenced to $t_{0}=50338.216$ (MJD). The errors are $90\%$ confidence.
A search of the $15-100$ keV HEXTE data for the pulsar, using the PCA
timing solution, failed to produce evidence of significant pulsations,
which is not surprising given the faintness of the source and the
presence of the bright confusing sources. The folded SGR 1900+14
pulsar lightcurve for three PCA energy ranges, using the above
timing parameters, is shown in Figure $2$. The pulsed fraction of
the SGR 1900+14 is not constrained by these data, due to the uncertain
x--ray flux from XTE J1906+09, 4U 1907+09, and the Galactic Ridge in
the {\it RXTE} bandpass.
\section{Discussion}
\label{discussion}
The $2-10$ keV SGR 1900+14 lightcurve obtained here is virtually
identical to the lightcurves obtained just before (\cite{hurley99b})
and just after (\cite{kouv99}) the commencement of the May 1999 active
period of the source. This indicates that the x--ray emitting geometry
is stable on timescales of years while the source is inactive. The
lightcurve appears to have multiple components which vary differently
with energy. There are three peaks in the $2-10$ keV lightcurve, with
a single relatively broad central peak surrounded by two narrower peaks.
The narrow peaks have harder spectra than the broad peak, as the narrow
peak emission dominates the emission from the broad peak above $10$ keV.
A simple explanation for the lightcurve morphology is that the pulsed
emission consists of different emission components arising from
different regions of the stellar surface. The narrow components may
be beamed emission from a collimated wind off of relatively small
hotspots, while the broader component could be more isotropic emission
from a larger and cooler area of the crust. The two narrow components
are greatly reduced in the pulsar lightcurves obtained just after the
giant flare of August 27, 1999 (\cite{kouv99}; \cite{murakami99}),
suggesting that the energy of the small hotspots may have been depleted
during the active period.
The observed temporal history of the SGR 1900+14 pulsar is shown in
Figure $3$. The additional timing parameters of the present observations
are important because they constrain the pulsar parameters long before
the source went into outburst. Although the temporal coverage is
incomplete, the {\it secular} spindown rate seems to change abruptly
sometime close to the initiation of bursting, at which point the
spindown continues steadily at an increased rate. These two different
spindown rates are denoted by the dotted lines in Figure $3$, which are
linear fits to the data before the outburst [up to and including the
first observation of Kouveliotou et al. (1999)] and the data during
and after the outburst [beginning with the first observation of
Kouveliotou et al. (1999) and ending with the Shitov (1999)
observation]. The third data point in Figure $3$, from Kouveliotou
et al. (1999), appears to be near the change point in the spindown
behavior because the period is consistent with the extrapolation of
the pre-outburst timing solution, yet the $\dot{P}$ value measured
during this observation is consistent with the outburst values. The
fit to the data taken during and after the outburst period yields a
value of $\dot{P}=(12.77\pm 0.01)\times 10^{-11}$ s s$^{-1}$ for the
mean spindown rate, and the corresponding pre-outburst value is
$\dot{P}=(6.126\pm 0.006)\times 10^{-11}$ s s$^{-1}$. Using these
mean $\dot{P}$ values, the mean inferred dipole field strengths before
and after the initiation of bursting would be $5.7\times 10^{14}$ G
and $8.2\times 10^{14}$ G, respectively, if the spindown were driven
by dipole radiation losses. These two values, which differ to a high
degree of significance, would imply an abrupt increase in the SGR
1900+14 magnetic field energy of more than $100\%$ around the time
the source started bursting, which is contrary to the predictions of
models in which the bursting is dissipating magnetic field energy.
This discrepancy clearly suggests that the SGR 1900+14 spindown is
not dominated by magnetic dipole radiation, and that the observed value
of $P\dot{P}$ provides no direct measurement of $B$, and no direct
evidence for a magnetar. Instead, the measured values of $P$ and
$\dot{P}$ suggest that the SGR spindown may be due to {\it winds},
if we take the pulsar age to be that of the associated (\cite{hurley99a}) supernova remnant G42.8+0.6. Assuming that the initial period of the
pulsar was much smaller than it is now, and that the braking index is
constant in time, the pulsar age $t_{age}=P/[(n-1)\dot{P}]$, where the
braking index $n$ is $3$ for pure dipole radiation but much less ($n\sim
1$) for spindown due to wind torques. Taking the estimated age of
G42.8+0.6 to be $\sim 10^{4}$ yr (\cite{vasisht94}, \cite{hurley96}),
we find that the braking index for SGR 1900+14 must be $\sim 1$, i.e.
$n = 1 + 0.16/(t_{age}/10^4\,\rm{yr})$, which indicates that the pulsar
spindown is dominated by winds. The remnant age would have to be an
order of magnitude smaller in order for the braking index to be consistent
with that of dipole radiation, and in addition such an age would require
an unreasonably large pulsar velocity of $\sim 2.5\times10^4$ km s$^{-1}$
for it to have traversed from the center of the remnant to its present
position, assuming a distance of 5 kpc (\cite{vasisht94}, \cite{hurley96}).
Thus the observations provide strong evidence that torques due to wind
emission, and not magnetic dipole torques, dominate the spindown dynamics
of SGR 1900+14.
The spindown behavior of SGR 1900+14 can be explained simply if we
assume that the spindown is due almost entirely to wind emission, as
was also considered by Kouveliotou et al. (1999). Possible mechanisms
for the generation of this wind include thermal radiation from hotspots
and Alfv\'{e}n wave emission (\cite{thompson98}). In this interpretation,
the SGR emits a robust wind of particles and fields, both during bursting
and quiescent intervals, which carries away angular momentum from the
star. The emission of a relativistic wind produces an exponential
spindown of the pulsar $\Omega(t)=\Omega_{0}\exp(-kt)$, where $k$ is
a constant parameterizing the rotational energy loss rate due to the
wind (\cite{thompson98}). Using this relation, and the values of $P$
and $\dot{P}$ from our observations, we obtain $k=\dot{P}/P\sim
2700^{-1}$ yr$^{-1}$. Given an age of $(1-2)\times 10^{4}$ yr for
G42.8+0.6, we obtain an initial pulsar spin period of $P_{0}\sim
3-120$ ms for SGR 1900+14, which is similar to the spin periods of
young isolated pulsars such as the Crab. This $P_{0}$ is most likely
an upper limit, given the likelihood of active periods (with higher
spindown rates) in the past.
As mentioned above, one scenario is that the spindown of SGR 1900+14
is due to Alfv\'{e}n wave emission, in which a stream of particles and
fields escape the star along magnetic field lines forced open by the
wind pressure (\cite{thompson98}). A supercritical magnetic field is
not required for this mechanism to explain the SGR 1900+14 spindown.
From Thompson \& Blaes (1998), the spindown constant is given by
\begin{equation}
k=1.5\times 10^{-11}{\left({B_{\ast}\over 3\times 10^{12}\,\rm{G}
}\right )}^{2}{\left({\delta B_{\ast}/B_{\ast}\over0.01}\right )}^{4/3}
\,\rm{Hz},
\end{equation}
where $B_{\ast}$ is the dipole field strength, $\delta B_{\ast}$
is the wave amplitude, and we have assumed a neutron star moment of
inertia and radius of $1.1\times 10^{45}$ g cm$^{2}$ and $10$ km,
respectively. This value of $k$ is comparable to the measured value
$k=\dot{P}/P\sim 10^{-11}$ Hz for SGR 1900+14, indicating that this
mechanism can explain the spindown of the SGR with conventional
($\sim 10^{12}$ G) field strengths, assuming that there is a
mechanism to continuously generate Alfv\'{e}n waves.
Even though a supercritical magnetic field on a global scale can not
account for the SGR pulsar spindown, such fields on much smaller
localized scales may nevertheless play an important role in the
bursting process. Since the wind torques initially operate to spin
down the neutron star crust, one might expect that if the core is not
rigidly coupled to the crust, then the core could be spinning slightly
faster and the resulting differential rotation could wind up any magnetic
field threading between the core and crust, building up large internal
magnetic field pressures. By analogy to the Sun, we might expect that
the growing pressure of the internal field is episodically released by
the surface break out of intense magnetic fields in localized regions,
similar to the appearance of sunspots, which have local fields of 10$^2$
to 10$^3$ times the average global surface field of the Sun. Such spots
of emerging magnetic flux (EMF) on a neutron star may thus contain
supercritical, or larger, localized fields, $B_s$ within radii $r_s$,
with total magnetic energies $> 3\times 10^{41}(B_s/B_{q})^2(r_s/1\,
\rm{km})^3$ erg, and they may be accompanied by comparable tectonic
stresses and heating from field diffusion in the crust. To contain
the giant flare of August 27, 1999, for example, a local field with
$B\sim B_{q}$ can contain the $3\times 10^{42}$ ergs of energy released (\cite{frail99}) within a bubble of radius $r_{s}\sim 2$ km, which is
a small fraction of the surface area of the star. The occurrence of such
EMF-spots could thus provide an episodic source of both magnetic and
tectonic-gravitational energy release, both thermal and nonthermal,
that power both the steady localized winds and the impulsive bursts
of SGRs, much as the sunspot fields are dissipated in winds, flares
and diffusion on the Sun. The solar analogy was also discussed by
Sturrock (1986) for Galactic gamma--ray bursts.
The SGR wind hypothesis can also explain other observed features of
the burst and quiescent emission from SGRs. If both the quiescent
x--ray emission and the spindown torque of SGR 1900+14 are due to
wind emission, the persistent x--ray flux and the spindown luminosity
should be correlated (this is not true of SGR 1806--20, because of
the surrounding plerion --- see below). Between the {\it ASCA}
observations of Hurley et al. (1999b) and Murakami et al. (1999),
the persistent x--ray flux of SGR 1900+14 increased by $(140\pm 20)
\%$. Using the appropriate mean $\dot{P}$ values from Figure $3$,
the spindown luminosity increased by $\sim 120\%$ over the same time
interval, which is consistent with the steady x--ray flux and spindown
arising from the wind.
The radio signature of SGR winds have been observed from SGRs 1900+14
(\cite{frail99}) and 1806--20 (\cite{kulkarni94}). In the latter case,
the SGR winds power a plerionic nebula with a total energy content
($\sim 10^{45}$ ergs) much greater than the energy given off in
a typical burst interval ($\sim 10^{43}$ ergs, \cite{kouv99}),
explaining the lack of variability seen from the SGR 1806--20 x--ray
and radio counterparts (\cite{sonobe94}; \cite{vasisht95}). In the
case of SGR 1900+14, a {\it transient} wind nebula from relativistic
particles injected during the giant flare of August 27, 1999
(\cite{hurley99c}) was observed by the VLA (\cite{frail99}). The
different radio properties of the SGR 1806--20 and SGR 1900+14
counterparts are probably due to the different external pressures
for the two sources, since SGR 1806--20 is still inside its high
pressure SNR while SGR 1900+14 is outside its associated supernova
remnant, where the confining pressure is relatively low. The weak
confining pressure of SGR 1900+14 inhibits the formation of a bright
plerion (\cite{frail99}). The observed nonthermal (photon index $\sim
2.2$: \cite{sonobe94}; \cite{hurley99b}) quiescent x--ray spectra of
the active SGR sources is characteristic of emission from a magnetized
wind (\cite{tavani94}). Finally, the burst spectra of SGRs can be
explained by the Compton upscattering of soft photons in a mildly
relativistic wind, without involving a supercritical stellar field (\cite{fat96}).
\acknowledgments
We thank Duane Gruber for suggesting improvements in the timing analysis.
This work was funded by NASA grant NAS5-30720.
|
\section{Introduction and statement of the results}
\subsection{Introduction}
\subsubsection{}
Scattering theory is a collection of methods and results for studying the continuous spectrum of operators.
Classical fields of application of scattering theory are the propagation of waves in mechanics, electro-dynamics and quantum mechanics. Often one considers a model that is close to some ideal case in which the wave equation can be solved explicitely. Scattering theory provides the tools for
comparison.
\subsubsection{}
In many cases it is possible to separate the time and space variables. In this case one can apply methods from stationary scattering theory. This branch of scattering theory investigates the generalized eigenfunctions contributing to the continuous spectrum. In good situations these eigenfunctions come in families which extend meromorphically. Often they are studied via a meromorphic continuation of the distribution kernel
of the resolvent of the spacial part of the operator.
\subsubsection{}
The problem of stationary scattering theory that is considered in the present paper has mainly an internal mathematical motivation. The model situation is a Laplace-type operator on a globally symmetric space
of negative curvature. Due to the presence of a large symmetry group it is possible to get an essentially complete description of the generalized eigenfunctions.
The real problem is to understand the generalized eigenfunctions on associated locally symmetric spaces.
The special case of quotients of the globally symmetric space by arithmetic subgroup groups has many applications in number theory and arithmetic. While arithmetic quotients have finite volume the main emphasis in the present work is on spaces of infinite volume. The question is, how far the interplay between global and local symmetries of the problem can be used in order to get a complete picture.
\subsubsection{}
There is lot of literature on the stationary scattering theory for the function Laplacian on spaces which become close to the global symmetric space asymptotically at infinity. The major part is devoted to the asymptotically hyperbolic case. The main results are meromorphic continuation of the resolvent kernel,
the scattering matrix and the parametrization of the relevant generalized eigenfunctions.
Most of this work is based on a fine study of the resolvent kernel. We refer to \cite{MR2153454} for one of the latest works and the discussion of the literature therein.
\subsubsection{}
In the more rigid case of a locally symmetric space one can obtain better results.
In this case one can approach stationary scattering theory by pushing the analysis to the boundary.
Using this method we obtained in \cite{MR1749869} the spectral decomposition of all locally invariant differential operators on vector bundles in the convex-cocompact case.
We refer to this paper for a review of the literature about the convex-cocompact case.
\subsubsection{}
The goal of the present paper is to generalize this method to geometrically finite spaces.
These are locally symmetric spaces of negative curvature which at infinity look like the symmetric space or a cusp. From the point of view of analysis the part at infinity which can be compared with the globally symmetric space is easy with the results of \cite{MR1749869} at hand.
The complications are due to the presence of cusps, in particular of those of non-maximal rank.
\subsubsection{}
Let $G$ be a real simple linear connected Lie group of real rank one.
It covers the connected component
of the group of isometries of the associated symmetric space $X$.
We consider a geometrically finite group $\Gamma\subset G$ which determines the locally symmetric space $\Gamma\backslash X$. For technical reasons we exclude the exceptional symmetric space $X={H\mathbb{O}}^2$
from our considerations.
The classical problems of meromorphic continuation of the Eisenstein series, the scattering matrix,
and their functional equations was previously
adressed by Guillop\'e \cite{guillope92} in the special case
that $X$ is the hyperbolic plane ${H\mathbb{R}}^2$ and by
Froese/Hislop/Perry \cite{MR1111571} for $X={H\mathbb{R}}^3$
and spherical Eisenstein series.
Cusps of non-maximal rank in the three-dimensional hyperbolic case have been investigated first in
\cite{MR1128217} via the resolvent. For the meromorphic continuation of the resolvent on
$\Gamma\backslash{H\mathbb{R}}^n$ see \cite{MR1710792},\cite{MR2209763}. All these papers are restricted to the case
of so-called rational cusps.
\subsubsection{}
The main result of the present paper is, in the geometrically finite case, the meromorphic continuation of the
Eisenstein series, i.e. of the families of generalized eigensections for all locally invariant differential operators on bundles, see Corollary \ref{main} below. Note that this result also includes the case of irrational cusps.
At the moment we are not able to obtain results, which are as complete as in the convex cocompact case.
In particular,
we are still quite far from showing a spectral decomposition.
\subsubsection{}
The class of geometrically finite discrete subgroups $\Gamma\subset G$
subsumes cocompact, convex-cocompact
groups, and subgroups of finite covolume as well as various kinds
of combinations of these cases with cusps of non-maximal rank.
The paper can be considered as a continuation
of \cite{MR1749869}, where the special case of a convex-cocompact
group $\Gamma$ is considered. We will frequently refer to this paper
for notations and conventions as well as for technical results.
\subsubsection{}
In the present paper the analysis takes place on the boundary $\partial X$ of
the symmetric space.
The group $\Gamma$ acts on vector bundles over this boundary.
As in \cite{MR1749869} the main players of the geometric
scattering theory are the push-down $\pi^\Gamma_*$ (average of sections over $\Gamma$),
the extension $ext^\Gamma$ (the adjoint ot the push-down), the restriction $res^\Gamma$, and the
scattering matrix $S^\Gamma$.
Our task is to construct these as families of maps between suitable function/distribution spaces
depending meromorphically on the spectral parameter.
Due to the presence of cusps of non-maximal rank the detailed description
of these spaces turns out to be quite difficult.
\subsection{Notion of geometrical finiteness}\label{geomf}
\subsubsection{}
In the present paper the symmetric space $X$ belongs to one of the series
of hyperbolic spaces ${H\mathbb{R}}^n$, ${H\mathbb{C}}^n$, and ${H\mathbb{H}}^n$
(real, complex, and quaternionic) of real dimension $n$, $2n$, $4n$.
We exclude the exceptional symmetric space of rank one, the Cayley hyperbolic plane ${H\mathbb{O}}^2$,
for technical reasons since we are going to employ the fact that
$X$ belongs to a series in several places.
\subsubsection{}
By $G$ we denote a real simple linear connected Lie group of real rank one
covering the connected component
of the group of isometries of $X$.
By $\partial X$ we denote the geodesic boundary
of $X$.
The union $\bar X:= X\cup \partial X$ has the structure of
a compact $G$-manifold with boundary. While $X$ parametrizes
the set of maximal compact subgroups the boundary $\partial X$
parametrizes the set of parabolic subgroups of $G$.
If $P\subset G$ is a parabolic subgroup, then we denote by
$\infty_P\in \partial X$ the unique fixed point of $P$.
Let $\Omega_P:=\partial X\setminus \infty_P$.
\subsubsection{}
Let $P\subset G$ be parabolic, and $N\subset P$ be its nil-radical.
Then we can write $P$ as an extension
\begin{equation}
\label{eepp}
0\rightarrow N\rightarrow P\stackrel{l}{\rightarrow} L\rightarrow 0\ .\end{equation}
Here $L$ is a reductive Lie group which is canonically
isomorphic to a product $MA$ of a compact group $M$ and a group $A\cong {\mathbb{R}}_+^*$.
\subsubsection{}
Let $\Gamma\subset G$ be a torsion-free discrete subgroup.
\begin{ddd}\label{weih111}
A parabolic subgroup $P\subset G$ is called $\Gamma$-cuspidal
if $U_P:=\Gamma\cap P$ is an infinite subgroup such that
$l(U_P)\subset L$ is precompact, i.e. $l(U_P)\subset M$.
\end{ddd}
Let $p$ denote the $\Gamma$-conjugacy class of the
$\Gamma$-cuspidal parabolic subgroup $P$. We call such classes cusps
and say that the cusp $p$ has full rank (as opposed to
smaller rank) if $U_P\backslash \Omega_P$ is compact for one (and hence for any) $P\in p$.
For each cusp $p$ and $P\in p$
we form the manifold with boundary $\bar Y_{U_P}:= U_P\backslash (X\cup \Omega_P)$.
\subsubsection{}\label{weih133}
The boundary $\partial X$ of $X$ has a $\Gamma$-equivariant decomposition
$\partial X=\Omega_\Gamma\cup\Lambda_\Gamma$
into a limit set $\Lambda_\Gamma$ and a domain of discontinuity
$\Omega_\Gamma$ (\cite{MR0336648}, Prop.~8.5). Let $\bar Y_\Gamma$ denote the manifold
with boundary $\bar Y_\Gamma:=\Gamma\backslash (X\cup \Omega_\Gamma)$.
\begin{ddd}\label{t67}
The group $\Gamma$ is called geometrically finite if the following conditions hold:
\begin{enumerate}
\item The set ${\mathcal{P}}_\Gamma$ of $\Gamma$-conjugacy classes of $\Gamma$-cuspidal parabolic subgroups
is finite.
\item There is a bijection ${\rm end}(\bar Y_\Gamma)\stackrel{\sim}{\rightarrow}
{\mathcal{P}}_\Gamma$, where ${\rm end}(\bar Y_\Gamma)$ denotes the set of ends of
the manifold with boundary $\bar Y_\Gamma$.
\item For all $p\in{\mathcal{P}}_\Gamma$ and $P\in p$ there is a representative
$\bar Y_p$ of the end corresponding to $p$ and an isometric embedding
$e_P: \bar Y_p \rightarrow \bar Y_{U_P}$ such that
its image $e_p(\bar Y_p)$ represents the end of $\bar Y_{U_P}$.
\end{enumerate}
\end{ddd}
By Bowditch \cite{bowditch951}, Corollary 6.3, this definition
is equivalent to the slightly weaker definition \cite{bowditch951}, F1
(which can be derived from \ref{t67} by replacing
``set ${\mathcal{P}}_\Gamma$ of $\Gamma$-conjugacy classes of $\Gamma$-cuspidal parabolic subgroups'' by ``set ${\rm end}(\bar Y_\Gamma)$ of ends of $\bar Y_\Gamma$" in 1., ``bijection" by ``map $c$" in 2. and ``all $p\in{\mathcal{P}}_\Gamma$" by
``all $p\in{\mathcal{P}}_\Gamma$ in the range of $c$" in 3.).
\subsubsection{}\label{weih132}
Let $Y_\Gamma:=\Gamma\backslash X$ denote the locally symmetric space associated to $\Gamma$.
By $\bar Y_0$ we denote the compact subset $\bar Y_\Gamma\setminus \bigcup_{p\in{\mathcal{P}}_\Gamma} \bar Y_{p}$
of $\bar Y_\Gamma$.
The boundary of $\bar Y_\Gamma$ is the manifold
$B_\Gamma:=\Gamma\backslash \Omega_\Gamma$.
Let ${\mathcal{P}}_\Gamma^<\subset {\mathcal{P}}_\Gamma$ denote the subset of cusps of smaller rank.
The decompositions of $\bar Y_\Gamma$ into a compact piece and its ends induces
a decomposition $B_\Gamma=B_0\cup \bigcup_{p\in {\mathcal{P}}^<_\Gamma} B_p$,
where $B_p:=B\cap \bar Y_p$, $p\in {\mathcal{P}}_\Gamma^<\cup\{0\}$.
\subsubsection{}\label{holle}
Set ${\mathcal{P}}_\Gamma^{max}:={\mathcal{P}}_\Gamma\setminus{\mathcal{P}}_\Gamma^<$. It will be convenient to fix
a set $\tilde{\mathcal{P}}$ of parabolic subgroups representing the elements of ${\mathcal{P}}_\Gamma$.
It comes with a partition $\tilde{\mathcal{P}}=\tilde{\mathcal{P}}^<\cup\tilde{\mathcal{P}}^{max}$.
\subsubsection{}\label{weih115}
Our analysis will require an additional assumption on the cusps.
By Lemma \ref{martin} this assumption is automatically
satisfied in the cases ${H\mathbb{R}}^n$, ${H\mathbb{C}}^n$, but by Lemma \ref{contrmartin}
it is non-trivial in the case ${H\mathbb{H}}^n$. We are now going to describe
this assumption in detail.
Let $P\subset G$ be parabolic. A Langlands decomposition
$P=MAN$ is the same as a split $s:MA\rightarrow P$ of the extension (\ref{eepp}),
where we identify $M$ and $A$ with their images $s(M)$ and $s(A)$.
Let $p$ be a cusp of $\Gamma$ and $P\in p$.
In Subsection \ref{cuspgeom} we will construct a Langlands
decomposition $P=MAN$, a discrete
subgroup $V\subset N$, and a homomorphism $m:N_V\rightarrow M$
from the Zariski closure $N_V$ of $V$ in $N$ such that
the group $U^0:=\{m(v)v|v\in V\}$ is a subgroup
of $U_P$ (see Def.~\ref{weih111}) of finite index.
\begin{ddd}\label{t799}
The cusp $p$ is called regular if
the Langlands decomposition of $P$ above can be adjusted such that
$N_V$ is invariant with respect to conjugation by $A$.
We call such a Langlands decomposition adapted.
\end{ddd}
\begin{ass}
In the remainder of the present paper we assume that
$\Gamma$ is geometrically finite and that all its cusps are
regular.
\end{ass}
\subsection{Twists and bundles}
\subsubsection{}\label{weih112}
We fix a parabolic subgroup $P$ of $G$ and write $\partial X=G/P$.
We further fix a Langlands decomposition $P=MAN$ and let ${\mathfrak{a}}$
denote the Lie algebra of $A$.
Let ${\mathfrak{n}}$ denote the Lie algebra of $N$, and let $\alpha\in{\mathfrak{a}}^*$
denote the short root of ${\mathfrak{a}}$ on ${\mathfrak{n}}$. We define
$\rho\in{\mathfrak{a}}^*$ by $\rho(H):=\frac12 {\mbox{\rm tr}}({\mbox{\rm ad}}(H)_{|{\mathfrak{n}}})$, $H\in{\mathfrak{a}}$.
\subsubsection{}\label{weih135}
If $\lambda\in {\aaaa_\C^\ast}$ and
$(\sigma,V_\sigma)$ is a representation of $M$, then we define the representation $(\sigma_\lambda,V_{\sigma_\lambda})$ of $P$
by $V_{\sigma_\lambda}:=V_\sigma$ and $\sigma_\lambda(man):=a^{\rho-\lambda}\sigma(m)$.
By $1$ we denote the trivial one-dimensional representation of $M$.
\subsubsection{}
If $(\theta,V_\theta)$ is a representation of $P$, then we define
the $G$-homogeneous bundle $V(\theta):=G\times_P V_\theta$
and denote by $\pi^\theta$ the representation of $G$
on spaces of sections of $V(\theta)$.
\subsubsection{}
If $({\varphi},V_{\varphi})$ is a finite-dimensional representation of $G$ (or $\Gamma$),
then we denote by $V(\theta,{\varphi})$ the tensor product
of $V(\theta)$ with the trivial bundle $\partial X\times V_{\varphi}$,
and by $\pi^{\theta,{\varphi}}$ the representation
of $G$ (or $\Gamma$) on spaces of sections of $V(\theta,\phi)$.
Note that we can identify
$$C^\infty(\partial X,V(\theta,{\varphi}))\cong C^\infty(\partial X,V(\theta))\otimes V_{\varphi},\quad \pi^{\theta,{\varphi}}=\pi^\theta\otimes{\varphi}\ .$$
\subsubsection{}
The representation ${\varphi}$ of $\Gamma$ is called twist.
Our analysis requires further assumptions on the twist going under the
name``admissible''.
Let $p$ be a cusp of $\Gamma$, $P\in p$, and $MAN$ be an adapted Langlands decomposition of $P$ (see Definition \ref{t799}). We define $M_U:=\overline{l(U_P)}$.
Then $P_U:=M_UN_V\subset P$ is a subgroup containing $U_P$.
Let $({\varphi},V_{\varphi})$
be a twist.
\begin{ddd}\label{weih134}
The twist $({\varphi},V_{\varphi})$ is called admissible at the cups $p$ if its restriction to $U_P$ extends to a continuous representation of $AP_U$ such that $A$ acts algebraically by
semisimple endomorphisms. A twist is called admissible
if it is admissible at all cusps of $\Gamma$.
\end{ddd}
Note that the algebraic functions on $A$ are linear combinations of $A\mapsto a^{n\alpha}$, $n\in {\mathbb{Z}}$
(see \ref{weih112} for the definition of $\alpha$).
\subsubsection{}
If $\Gamma$ has cusps, then the condition of admissibility excludes most unitary representations of $\Gamma$. Examples of admissible twists are restrictions
of finite-dimensional representations of $G$ to $\Gamma$.
As explained in Subsection \ref{ttw} twists are used
as a technical device.
\subsubsection{}
If $(\gamma,V_\gamma)$ is a representation of $K$, then we can form the
homogeneous bundle $V(\gamma):=G\times_KV_\gamma$ over $X$.
For a twist ${\varphi}$ let $V(\gamma,{\varphi})$ denote the product
$V(\gamma)\otimes V_{\varphi}$. Furthermore let $V_\Gamma(\gamma,{\varphi}):=\Gamma\backslash V(\gamma,{\varphi})$
denote the corresponding bundle over $Y$.
\subsection{Exponents}
\subsubsection{}
The main numerical invariant associated to a discrete subgroup $\Gamma\subset G$
is its exponent $\delta_\Gamma\in {\mathfrak{a}}^*$. Note that each $g\in G$ can be written
as $g=ka_gh\in KA_+K$, where $a_g$ is uniquely determined. Here $A_+:=\{a\in A\mid a^\alpha\ge 1\}$.
\begin{ddd}
The exponent $\delta_\Gamma\in{\mathfrak{a}}^*$ is defined as the infimum
of the set $$\{\nu\in{\mathfrak{a}}^*\:|\:\sum_{g\in\Gamma} a_g^{-\nu-\rho}<\infty\}\ .$$
\end{ddd}
If follows from the discreteness of $\Gamma$ that
$\delta_\Gamma\le\rho$.
\subsubsection{}
The critical exponent $\delta_\Gamma$ has been extensively studied, in particular by Patterson \cite{patterson762}, Sullivan \cite{sullivan79}, \cite{MR766265}, and Corlette \cite{corlette90}, Corlette and Iozzi \cite{corletteiozzi99}. From these papers we know that $\delta_\Gamma\in [-\rho,\rho]$, if $\Gamma$ is infinite. Moreover, we have $\delta_\Gamma=\rho$ if and only if $\Gamma$ has finite covolume.
If $\Lambda_\Gamma$ contains at least $2$ points, then $\delta_\Gamma+\rho=\dim_H(\Lambda)\alpha$, where $\dim_H(\Lambda)$ denotes the Hausdorff dimension of the limit set with respect to the natural class of sub-Riemannian metrics on $\partial X$. If $Y=\Gamma\backslash X$ is an infinite volume quotient of a quaternionic hyperbolic space or the
Cayley hyperbolic plane, then $\delta_\Gamma$ can not be arbitrary close to
$\rho$. In these cases we have $\delta_\Gamma\le (2n-1)\alpha$ and $\delta_\Gamma\le 5\alpha$, respectively \cite{corlette90},\cite{corletteiozzi99}.
\subsubsection{}
Let ${\varphi}$ be a twist.
\begin{ddd} We define the exponent $\delta_{\varphi}\in{\mathfrak{a}}^*$
of ${\varphi}$ as the infimum of the set $$\{\nu\in{\mathfrak{a}}\:|\:\sup_{g\in\Gamma} a_g^{-\nu} \|{\varphi}(g)\|<\infty\}\ ,$$
where $\|.\|$ denotes any norm on ${\mathrm{End}}(V_{\varphi})$.
\end{ddd}
\subsection{Description of the main results}
\subsubsection{}
Let $\Gamma\subset G$ be a discrete, torsion-free, geometrically finite
subgroup such that all its cusps are regular. Furthermore let
${\varphi}$ be an admissible twist. Let $B_\Gamma:=\Gamma\backslash \Omega_\Gamma$ and
$V_{B_\Gamma}(\sigma_\lambda,{\varphi}):=\Gamma\backslash V(\sigma_\lambda,{\varphi})$.
Under the name push-down we subsume several constructions related
to the average of elements of $C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$
with respect to $\Gamma$. It is a crucial matter to construct
a space $B_\Gamma(\sigma_\lambda,{\varphi})$ which contains the result
of the average. If $\lambda\in{\aaaa_\C^\ast}$ varies, then these spaces
assemble as a projective limit of locally trivial bundles
of Fr\'echet spaces in the sense of Subsection
\ref{llim}. Therefore we can speak of meromorphic families of continuous maps
from and to these spaces.
\subsubsection{}\label{weih123}
We define the bundle $B_\Gamma(\sigma_\lambda,{\varphi})$
as a direct sum
$$B_\Gamma(\sigma_\lambda,{\varphi}):=B_\Gamma(\sigma_\lambda,{\varphi})_1 \oplus {R}_{\Gamma}(\sigma_\lambda,{\varphi})_{max}\ ,$$
where the first component satisfies
$$C_c^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))\subset B_\Gamma(\sigma_\lambda,{\varphi})_1\subset C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$
(equality occurs if all cusps have full rank).
The second component is associated to the cusps of full rank and is finite-dimensional.
\subsubsection{}\label{neuj913}
The elements of $B_\Gamma(\sigma_\lambda,{\varphi})_1$ are characterized as smooth
sections of $V_{B_\Gamma}(\sigma_\lambda,{\varphi})$ with a certain asymptotic expansion near the cusps.
We postpone the detailed description of the asymptotics until Subsection \ref{asz}.
The bundle ${R}_{\Gamma}(\sigma_\lambda,{\varphi})_{max}$ has a further decomposition
$${R}_{\Gamma}(\sigma_\lambda,{\varphi})_{max}=\bigoplus_{P\in\tilde{\mathcal{P}}^{max}} {R}_{U_P}(\sigma_\lambda,{\varphi})\ .$$
Recall from \ref{holle} that $\tilde {\mathcal{P}}^{max}$ is in bijection with the set of cusps of full rank. In order to define
${R}_{U_P}(\sigma_\lambda,{\varphi})$ we first consider the sheaf
${\mathcal{E}}_{\infty_P}(\tilde\sigma,\tilde{\varphi})$ on ${\aaaa_\C^\ast}$ of holomorphic families $\phi_\nu$, $\nu\in{\aaaa_\C^\ast}$, of $U_P$-invariant distribution sections
of $V(\tilde\sigma_{\nu},\tilde{\varphi})$ supported in $\infty_P$.
Here $\tilde\sigma,\tilde{\varphi}$ are the dual representations to
$\sigma,{\varphi}$. The sheaf
${\mathcal{E}}_{\infty_P}(\tilde\sigma,\tilde{\varphi})$ is the sheaf of holomorphic sections of a trivial finite-dimensional holomorphic
vector bundle over ${\aaaa_\C^\ast}$, and we denote by $E_{\infty_P}(\tilde\sigma_\lambda,\tilde{\varphi})$
the fibre of this bundle over $\lambda\in{\aaaa_\C^\ast}$
(see Lemma \ref{vermaolbrich} for an alternative description of the space
$E_{\infty_P}(\tilde\sigma_\lambda,\tilde{\varphi})$ and \ref{weih121} for the finite-dimensionality).
Then we define
$${R}_{U_P}(\sigma_\lambda,{\varphi}):= E_{\infty_P}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*\ .$$
\subsubsection{}
We can now state the definition of the push-down.
\begin{ddd}\label{pushdowndef}
For $f\in C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$
we define the push-down $\pi^\Gamma_*(f)\in B_\Gamma(\sigma_\lambda,{\varphi})$
as the direct sum of $\pi^\Gamma_*(f)_1\in B_\Gamma(\sigma_\lambda,{\varphi})_1$
and $\pi^\Gamma_*(f)_2\in {R}_{\Gamma}(\sigma_\lambda,{\varphi})_{max}$.
Here $\pi^\Gamma_*(f)_1$ is given by
$\pi^\Gamma_*(f)_1:=\sum_{g\in\Gamma} \pi^{\sigma_\lambda,{\varphi}}(g) f_{|\Omega_\Gamma}$
(if the sum converges). The component
$\pi^\Gamma_*(f)_{2,P}\in {R}_{U_P}(\sigma_\lambda,{\varphi})$ for $P\in\tilde{\mathcal{P}}^{max}$ is defined
by the condition that
$$\langle \phi,\pi^\Gamma_*(f)_{2,P}\rangle = \sum_{[g]\in\Gamma/U_P}\langle \pi^{\tilde\sigma_{-\lambda},\tilde{\varphi}}(g)\phi,f \rangle$$
for all $\phi\in E_{\infty_P}(\tilde\sigma_{-\lambda},\tilde{\varphi})$
(if the sum converges).
\end{ddd}
\subsubsection{}
In \ref{pushdowndef} the push-down is defined if certain sums converge. The first part of the following theorem gives a range of $\lambda\in {\aaaa_\C^\ast}$ for which these convergence conditions are satisfied.
The second part asserts its meromorphic continuation to all of ${\aaaa_\C^\ast}$.
\begin{theorem}\label{t119}
If $f\in C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$, then the push-down
$\pi^\Gamma_*(f)\in C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$ converges for ${\rm Re }(\lambda)<-\delta_\Gamma-\delta_{\varphi}$.
The push-down induces a meromorphic family on ${\aaaa_\C^\ast}$ of continuous maps
$$\pi^\Gamma_*:C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
B_\Gamma(\sigma_\lambda,{\varphi})$$ with finite-dimensional singularities.
\end{theorem}
We will first prove the part of the theorem asserting the convergence of the
push-down for ${\rm Re }(\lambda)<-\delta_\Gamma-\delta_{\varphi}$. Then we consider
the adjoint of the push-down, the extension $ext^\Gamma$,
and obtain the meromorphic continuation of the latter (see Theorem
\ref{t110}).
The remainder of Theorem \ref{t119} then follows from this result by duality.
\subsubsection{}\label{weih126}
By $D_\Gamma(\sigma_\lambda,{\varphi})$ we denote the topological dual space
of $B_\Gamma(\tilde \sigma_{-\lambda},\tilde{\varphi})$.
It is a dual Fr\'echet and Montel space.
By the latter property it is reflexive so that $$D_\Gamma(\sigma_\lambda,{\varphi})^\prime\cong
B_\Gamma(\tilde \sigma_{-\lambda},\tilde{\varphi})\ .$$
Varying $\lambda\in{\aaaa_\C^\ast}$ the spaces $D_\Gamma(\sigma_\lambda,{\varphi})$
form a direct limit of locally trivial holomorphic bundles of dual Fr\'echet
spaces in the sense of Subsection \ref{llim}.
\begin{ddd}
For ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$
the extension
$$ext^\Gamma:D_\Gamma(\sigma_\lambda,{\varphi})\rightarrow
C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$$
is defined to be the adjoint of
$\pi^\Gamma_*:C^\infty(\partial X,V(\tilde \sigma_{-\lambda},\tilde{\varphi}))\rightarrow B_\Gamma(\tilde \sigma_{-\lambda},\tilde{\varphi})$.
\end{ddd}
\begin{theorem}\label{t110}
The extension induces a meromorphic family of continuous maps on ${\aaaa_\C^\ast}$
$$ext^\Gamma:D_\Gamma(\sigma_\lambda,{\varphi})\rightarrow
C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$$
with finite-dimensional singularities and values in ${}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$.
\end{theorem}
\subsubsection{}
The meromorphic continuation of the extension is the main goal
of the present paper. We obtain the continuation in close connection
with the meromorphic continuation of the scattering matrix
which will be defined below. In order to define this scattering matrix
we need the restriction map $res^\Gamma$ which is a left-inverse of the
extension. Due to the presence of cusps the definition of
the restriction map is more complicated than in \cite{MR1749869}.
\subsubsection{}
We first recall the definition of $res^\Gamma$ in the case
of convex-cocompact groups $\Gamma$ (i.e. in the case without cusps, compare Sec.~4 of \cite{MR1749869}). In this case
$$res^\Gamma:{}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
C^{-\infty}(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$
is given by the composition of the restriction of distributions
to $\Omega_\Gamma$
and the identification ${}^\Gamma C^{-\infty}(\Omega_\Gamma,V(\sigma_\lambda,{\varphi}))\cong C^{-\infty}(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$.
For the purpose of the present paper we employ a different
description.
\subsubsection{}
Let us still assume that $\Gamma$ is convex-cocompact.
We choose a cut-off function
$\chi^\Gamma\in C_c^\infty(\Omega_\Gamma)$ such that
$\sum_{g\in\Gamma} g^* \chi^\Gamma=1$.
Then we define $$\pi^{*}:
C^\infty(B_\Gamma,V_{B_\Gamma}(\tilde\sigma_{-\lambda},\tilde{\varphi}))
\rightarrow C^\infty(\partial X,V(\tilde\sigma_{-\lambda},\tilde{\varphi}))$$ by
$\pi^*(f)=\chi^\Gamma f$ (in order to understand the right-hand side properly one must identify sections of bundles over $B_\Gamma$ with $\Gamma$-equivariant sections on $\Omega_\Gamma$).
We define $$\widetilde{res}:C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))
\rightarrow C^{-\infty}(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$
to be the adjoint of $\pi^*$. Then
$res^\Gamma$ coincides with the restriction of $\widetilde{res}$
to the subspace ${}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$.
While $\widetilde{res}$ depends on the choice of $\chi^\Gamma$,
the restriction $res^\Gamma$ is independent of choices.
\subsubsection{}
If $\Gamma$ has cusps of smaller rank, then we can not assume
that $\chi^\Gamma$ has compact support. The definition of $\pi^*$ breaks down.
Our way arround that problem is as follows.
For each $k\in{\mathbb {N}}_0$ we consider the
Banach spaces $C^k(\partial X,V(\tilde\sigma_{-\lambda},\tilde{\varphi}))$
and define a meromorphic family of maps
$\pi^*_k:B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})\rightarrow
C^k(\partial X,V(\tilde\sigma_{-\lambda},\tilde{\varphi}))$.
The map $\pi^*_k$ is a right-inverse of the push-down $\pi^\Gamma_*$,
and its definition is similar in spirit to that
of $\pi^*$ above, but the details are more complicated
since we have to take into account the asymptotic expansions of the elements of
$B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})$ near cusps.
One problem is that we can only deal with a finite number (depending on $k$) of these terms at a time in order to define $\pi_k^*$. The situation is similar to the problem of the construction
of smooth functions with given Taylor series at a given point. It is impossible to construct a continuous map from the space of formal power series to smooth functions which is right-inverse to the surjective map which takes the Taylor series.
Back to the definition of the restriction, let $C^{-k}(\partial X,V(\sigma_\lambda,{\varphi}))$ denote the
distributions of order $k$, i.e. the topological dual of
$C^k(\partial X,V(\tilde\sigma_{-\lambda},\tilde{\varphi}))$,
and define
$res_k^\Gamma:C^{-k}(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
D_\Gamma(\sigma_\lambda,{\varphi})$ as the adjoint of $\pi^*_k$.
The collection of maps $res_k^\Gamma$ play the role of the map
$\widetilde{res}$ above. These maps depend on choices
and are, in particular, not compatible if we vary $k$.
\subsubsection{}
Nevertheless, we have the following uniqueness property.
If $f_\nu\in {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\nu,{\varphi}))$,
$\nu\in{\aaaa_\C^\ast}$, is a meromorphic family, and $W\subset {\aaaa_\C^\ast}$
is compact, then for sufficiently large $k$
the meromorphic family $res^\Gamma_k(f_\nu)$, $\nu\in W$, is well-defined. In fact,
due to the compactness of $W$ the order of $f_\nu$ as a distribution is uniformly bounded for $\nu\in W$.
In \ref{neuj3001} we will see that
$res^\Gamma_k(f_\nu)$ is independent of the choices.
\subsubsection{}
In general the restriction maps $res^\Gamma_k$ may have poles.
Moreover, not every element of ${}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$
can be written as evaluation of a holomorphic family $f_\nu\in {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\nu,{\varphi}))$ defined near $\lambda$. But for generic $\lambda\in {\aaaa_\C^\ast}$ the restriction
map $res^\Gamma_k$ is regular and every element $f\in {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$ extends to a family (in the present paper we will prove a weaker statement only, which suffices for our purposes).
In this case
$res^\Gamma_k(f)$ is well-defined independent of the choice of (the sufficiently large) $k$,
and we can omit the subscript $k$ and write
$res^\Gamma(f)$ for $res^\Gamma_k(f)$.
We have the identity
$$res^\Gamma\circ ext^\Gamma={\mathrm {id}}_{D_\Gamma(\sigma_\lambda,{\varphi})}\ ,$$
which holds true for generic $\lambda\in {\aaaa_\C^\ast}$ or as an identity of maps defined on meromorphic families
(see Lemma \ref{weih114}).
\subsubsection{}
If $f\in {}^\Gamma C^{-k}(\partial X,V(\sigma_\lambda,{\varphi}))$, $k\in{\mathbb {N}}_0$,
then $f_{|\Omega_\Gamma}$ is a $\Gamma$-invariant distribution
on $\Omega_\Gamma$, hence can be considered as an element of
$C^{-\infty}(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$. If $$i:D_\Gamma(\sigma_\lambda,{\varphi})
\rightarrow C^{-\infty}(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$
is the natural map
(adjoint to the inclusion $C_c^\infty(B,V_B(\tilde\sigma_{-\lambda},\tilde{\varphi}))
\hookrightarrow B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})$),
then $i\circ res^\Gamma(f)$ is defined (even though
$res^\Gamma$ may have a pole at $\lambda$)
and coincides with $f_{|\Omega_\Gamma}$.
In particular, the condition
$i\circ res^\Gamma(f)=0$ is equivalent to ${\mathrm{supp}}(f)\subset \Lambda_\Gamma$.
If $\Gamma$ has cusps, then $i$ is not injective.
In this case the condition ${\mathrm{supp}}(f)\in \Lambda_\Gamma$ does not imply
that $res^\Gamma(f)=0$ (provided the latter is defined).
\subsubsection{}
Vanishing of $res^\Gamma(f)$ is a rather strong condition
and should play the role of the condition ${\mathrm{supp}}(f)\subset \Lambda_\Gamma$ in the convex-cocompact case which was successfully exploited in \cite{MR1749869} and \cite{MR1689342}.
One application of this condition is the following result
which is employed in proving the functional equation of the scattering matrix in the domain of convergence.
\begin{theorem}[Corollary \ref{wieder}]\label{micro}
Assume that ${\rm Re }(\lambda)>\max\left(\{\delta_\Gamma\}\cup\{\rho_{U_P}\mid P\in\tilde{\mathcal{P}}\}\right)+\delta_{\varphi}$. If
$f_\mu\in {}^\Gamma C^{-\infty}(\partial X,V(1_\mu,{\varphi}))$
is a germ of a meromorphic family at $\lambda$, then
$ext^\Gamma\circ res^\Gamma(f_\mu) = f_\mu$.
\end{theorem}
Further applications of the condition $res^\Gamma(f)=0$
are contained in \cite{MR1926489} (non-existence of such $f$ for ${\rm Re }(\lambda)$ large) and \cite{math.DG/0103144} (estimates of the regularity of $f$). In fact, the proof of
Theorem \ref{micro} is based on the main result of \cite{MR1926489}.
\subsubsection{}
We now turn to the scattering matrix. We assume that $\sigma$ is either
a Weyl-invariant irreducible representation of $M$ or
of the form $\sigma^\prime\oplus(\sigma^\prime)^w$,
where $\sigma^\prime$ is irreducible and not Weyl-invariant,
and $(\sigma^\prime)^w(m):=\sigma^\prime(w^{-1}mw)$
is the Weyl-conjugate representation of $\sigma^\prime$.
Here $w\in N_K(A)$ is a representative of the non-trivial
element of the Weyl group $W=N_K(A)/M\cong {\mathbb{Z}}/2{\mathbb{Z}}$.
\subsubsection{}\label{gaa101}
Our starting point
is the scattering matrix
$$S^{\{1\}}_\lambda=J_\lambda:C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi})) \rightarrow
C^{-\infty}(\partial X,V(\sigma_{-\lambda},{\varphi}))$$ associated to the trivial group $\{1\}$. In representation theory it firms under the name
Knapp-Stein intertwining operator.
Here we employ a suitably normalized version as in the paper
\cite{MR1749869}, Sec. 5, to which we refer for further details.
Let $I_{\mathfrak{a}}\subset{\mathfrak{a}}^*$ be the ${\mathbb{Z}}$-module spanned by
$\alpha$ if $2\alpha$ is a root of $({\mathfrak{g}},{\mathfrak{a}})$, and by
$\frac{1}{2}{\alpha}$ otherwise.
The operators $J_\lambda$ form a meromorphic family
of continuous maps with singularities in $I_{\mathfrak{a}}$
and satisfy the functional equation $J_\lambda\circ J_{-\lambda}={\mathrm {id}}$.
\subsubsection{}
If $\Gamma$ is non-trivial, then we obtain the
scattering matrix $S_\lambda^\Gamma$ from $J_\lambda$
using restriction and extension.
\begin{ddd}
For ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$ we define the scattering matrix
$$S^\Gamma_\lambda:D_\Gamma(\sigma_\lambda,{\varphi})\rightarrow D_\Gamma(\sigma_{-\lambda},{\varphi})$$
as the meromorphic
family of continuous maps
$$S^\Gamma_\lambda:=res^\Gamma\circ J_\lambda\circ ext^\Gamma\ .$$
\end{ddd}
\begin{theorem}\label{t113}
The scattering matrix has a meromorphic
continuation to all of ${\aaaa_\C^\ast}$. It
satisfies the functional equation $S^\Gamma_\lambda\circ S^\Gamma_{-\lambda}={\mathrm {id}}$ and
the relation $J_\lambda\circ ext^\Gamma=ext^\Gamma\circ S^\Gamma_\lambda$.
\end{theorem}
As noted above this theorem is proved in a multistep
procedure which involves the meromorphic continuation of $ext^\Gamma$
at the same time. The basic ideas are adapted
from \cite{MR1749869}. The non-compactness
of $B_\Gamma$ due to the presence of cusps of non-maximal
rank is responsible for the various complications which
have to be resolved on the way.
\subsubsection{}
The following application to the Eisenstein series is
an easy consequence of the preceding
results and can be derived exactly as in \cite{MR1749869},
Cor.~10.2.
Let $\gamma$ be a finite-dimensional representation of $K$ and $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$.
Then we have a holomorphic family of $G$-equivariant maps
$$P^T_\lambda: C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
C^\infty(X,V(\gamma,{\varphi}))$$
which are called Poisson transformations.
We refer to \cite{MR1749869}, Def. 4.8, for a definition of the Poisson
transformation and to \cite{MR1749869}, Sec. 6, for a discussion of its properties.
\begin{ddd}\label{neuj1001}
If $\lambda\in{\aaaa_\C^\ast}$, $f\in D_\Gamma(\sigma_\lambda,{\varphi})$,
$ext^\Gamma(f)$ is regular,
and $T\in {\mbox{\rm Hom}}_M(V_\sigma,V_\gamma)$, then
we define the Eisenstein series
$E(\lambda,f,T)\in C^\infty(Y,V_Y(\gamma,{\varphi}))$
by $P^T_\lambda\circ ext^\Gamma(f)$.
\end{ddd}
In the special case that $\sigma$, $\gamma$, and ${\varphi}$ are trivial, $T$ is
the identity, ${\rm Re }(\lambda)>\delta_\Gamma$, and $f=\delta_b$ is the
delta-distribution located at $b\in B_\Gamma$, this definition
coincides with the classical definition of the Eisenstein series
as the $\Gamma$-average of the Poisson kernel, i.e. the
integal kernel of $P^{{\mathrm {id}}}_\lambda$.
Let $c_\sigma$ and $c_\gamma$ be the $c$-functions introduced in \cite{MR1749869}, Sec. 5.
We further employ the notation $(c_\gamma(\lambda) T)^w$ as introduced in the same section of \cite{MR1749869}.
\begin{kor}\label{main}
The Eisenstein series forms a meromorphic family of continuous maps
with finite-dimensional singularities
$$E(\lambda,.,T): D_\Gamma(\sigma_\lambda,{\varphi}) \rightarrow C^\infty(Y,V_Y(\gamma,{\varphi}))\ .$$
It satisfies the functional equation
$$E(\lambda,S_{-\lambda} f,T)=E(-\lambda,f,(\frac{c_\gamma(\lambda)}{c_\sigma(\lambda)}T)^w)\ .$$
\end{kor}
\section{Some machinery}
\subsection{Limits of bundles}\label{llim}
\subsubsection{}
In the present paper a main issue is the construction of
families of topological vector spaces $V_\lambda$, $\lambda\in{\mathbb {C}}$, and
the study of holomorphic (meromorphic) families of vectors
$f_\lambda\in V_\lambda$ or holomorphic (meromorphic) families of
homomorphisms $h_\lambda\in {\mbox{\rm Hom}}(V_\lambda,W_\lambda)$
between such families of spaces. In order to make the notion
of a holomorphic family precise we must relate the spaces $V_\lambda$
for neighbouring $\lambda$ in some holomorphic manner.
One way to do this is to equip the family of spaces with the
structure of a locally trivial holomorphic bundle of topological vector
spaces. If we fix two local trivializations $U_0\times W_0$,
$U_1\times W_1$ of the family $V_\lambda$ with non-trivial overlap,
then the transition function is a holomorphic map from $U_0\cap U_1$ to
${\mbox{\rm Hom}}(W_0,W_1)$ with values in the subspace of isomorphisms.
\subsubsection{}
All topological vector spaces in the present paper will be locally convex.
Spaces of homomorphisms between topological vector spaces
will always be equipped with the topology of uniform convergence
on bounded sets. We refer to \cite{MR1689342}, Sec.2.2
for more details.
If we have two holomorphic families
$h_\lambda\in{\mbox{\rm Hom}}(W_0,W_1)$ and $g_\lambda\in{\mbox{\rm Hom}}(W_1,W_2)$,
then we can consider the composition $g_\lambda\circ h_\lambda\in {\mbox{\rm Hom}}(W_0,W_2)$.
This composition is again holomorphic (see \cite{gloeckner}).
If the evaluation $W_0\to W^{\prime\prime}_0$ is continuous, then the adjoint
$h_\lambda^\prime\in {\mbox{\rm Hom}}(W_1^\prime,W_0^\prime)$ is holomorphic, too (see again \cite{gloeckner}).
In order to show that the composition of two holomorphic families of maps
is again holomorphic in \cite{MR1689342}, Sec.2.2, we assumed that $W_0$ is a Montel space. The discussion of \cite{gloeckner} shows that this assumption can be omitted.
\subsubsection{}
In this paper we do not show that the spaces $B_\Gamma(\sigma_\lambda,{\varphi})$
($D_\Gamma(\sigma_\lambda,{\varphi})$) form locally trivial bundles of Fr\'echet
(dual Fr\'echet) spaces. These families of spaces arise instead as projective (symbol $\lim$) and inductive limits (symbol $\mathrm{colim}$) of locally trivial bundles, respectively.
E.g. the bundle $B_U(\sigma_\lambda,{\varphi})$ comes as a limit $\lim B_{\Gamma,k}(\sigma_\lambda,{\varphi})$
of trivial bundles $B_{\Gamma,k}(\sigma_\lambda,{\varphi})$,
$k\in {\mathbb {N}}_0$. The connecting maps (in the example $B_{\Gamma,k}(\sigma_\lambda,{\varphi})\to B_{\Gamma,k+1}(\sigma_\lambda,{\varphi})$) of the diagrams are holomorphic families of continuous maps
which are injective and have dense range (resp. are injective).
But they are not compatible with the trivializations.
\subsubsection{}\label{weih211}
We must extend the notions of a holomorphic or meromorphic
family of continuous of maps between such limits of locally trivial holomorphic bundles.
Let $$\dots\subset E_{n+1}\subset E_{n} \subset\dots\ ,$$
$$\dots\subset F_{n+1}\subset F_{n} \subset\dots$$
be decreasing families of locally trivial bundles of Fr\'echet spaces
defined over some open subset $U\subset {\mathbb {C}}$
such that all inclusion maps are holomorphic.
Let $E:=\lim E_n$, $F:=\lim F_n$.
\begin{ddd}
A family of maps $\phi_z:E_z\rightarrow F_z$, $z\in U$,
is called holomorphic (meromorphic) if
for each $x\in U$ and $n\in {\mathbb {N}}_0$ there exists
a neighbourhood $U_{x,n}\subset U$ of $x$, $m(n)\in{\mathbb {N}}$, and
a holomorphic (meromorphic) bundle map
$\Phi_n: (E_{m(n)})_{|U_{x,n}}\rightarrow (F_{n})_{|U_{x,n}}$ such that
$\phi_y$ is the restriction of $\Phi_n(y)$ to $E_y$
for all $y\in U_{n,x}$.
\end{ddd}
The composition of two holomorphic families is again a holomorphic family
(compare \cite{MR1749869}, Subsection 2.2).
\subsubsection{}
We now consider the dual situation.
Let $$\dots \rightarrow E^\prime_n\rightarrow E^\prime_{n+1}\rightarrow\dots\ ,$$
$$\dots \rightarrow F^\prime_n\rightarrow F^\prime_{n+1}\rightarrow\dots$$
be direct systems of holomorphic locally trivial bundles
of dual Fr\'echet spaces defined over some open
subset $U\subset {\mathbb {C}}$. Let $E^\prime:=\mathrm{colim} E^\prime_n$,
$F^\prime:=\mathrm{colim} F^\prime_n$.
\begin{ddd}
A family of maps $\phi_z:F_z^\prime\rightarrow E_z^\prime$, $z\in U$,
is called holomorphic (meromorphic) if
for each $x\in U$ and $n\in{\mathbb {N}}_0$
there exists
a neighbourhood $U_{x,n}\subset U$ of $x$, $m(n)\in{\mathbb {N}}$,
and a holomorphic (meromorphic) bundle map
$\Phi_n^\prime: (F^\prime_{n})_{|U_{x,n}}\rightarrow (E^\prime_{m(n)})_{|U_{x,n}}$
such that for each $y\in U_{x,n}$
the restriction of $\phi^\prime_y$ to $(F^\prime_n)_y$
is the composition of $(\Phi_n^\prime)_y$ with the
natural map $(E^\prime_{m(n)})_y\rightarrow E^\prime_y$ .
\end{ddd}
Again the composition of two such families is a holomorphic (meromorphic) family. The adjoint of a holomorphic family is again a holomorphic (meromorphic) family.
\subsection{Embedding}\label{embedd}
\subsubsection{}
The symmetric spaces
$X$ considered in the present paper belong to a series of symmetric spaces.
Let $X^n$ denote the $n$'th space of the series.
We will use the superscript
${}^n$ as a decoration of symbols for other objects in order to indicate
that they are associated to $X^n$. For a number of arguments we need
that $\delta_\Gamma^n$ is sufficiently negative. Note that
$\delta_\Gamma^n\to-\infty$ as $n\to\infty$. E.g., we will first obtain
a meromorphic continuation of $ext^{\Gamma,n+k}$ for sufficiently
large $k$, and then use the propositions below in order to conclude that
$ext^{\Gamma,n}$ has a meromorphic continuation, too.
\subsubsection{}
The main point of this subsection is to explain that the concept of embedding is compatible with the function and distribution spaces introduced in \ref{weih123} and \ref{weih126}. Furthermore we need compatibility with the push-down and extension maps. Note that the concept of embedding is only applied to the spherical case $\sigma=1$.
\subsubsection{}\label{neuj700}
We have embeddings $X^n\hookrightarrow X^{n+1}$ and
$i:\partial X^n\hookrightarrow \partial X^{n+1}$.
In order to have compatible embeddings $G^n\hookrightarrow G^{n+1}$
of the groups we assume
at this point that $G^n$ is one of $$\{Spin(1,n),SO(1,n)_0,SU(1,n),Sp(1,n)\}\ .$$
If we realize the last three groups of the list
using ${\mathbb{F}}$-valued matrices, ${\mathbb{F}}\in\{{\mathbb{R}},{\mathbb {C}},{\mathbb{H}}\}$, then the
embedding $G^n\hookrightarrow G^{n+1}$ is the usual embedding
into the left upper corner. We have compatible Iwasawa decompositions
such that $A=A^n=A^{n+1}$, and in particular $P^n\hookrightarrow P^{n+1}$.
Let $\zeta$ be $\alpha/2$, $\alpha$, and $2\alpha$ in the cases
$X={H\mathbb{R}}^n$, $X={H\mathbb{C}}^n$, and $X={H\mathbb{H}}^n$, respectively.
Then $\zeta=\rho^{n+1}-\rho^n$, and we have
$(V_{1^{n+1}_\lambda})_{|P^n}=V_{1^n_{\lambda-\zeta}}$, and
hence $V(1^{n+1}_\lambda,{\varphi})_{|\partial X^n}=V(1^n_{\lambda-\zeta},{\varphi})$.
\subsubsection{}\label{weih301}
Let
$$i^*:C^\infty(\partial X^{n+1},V(1^{n+1}_\lambda,{\varphi}))\rightarrow
C^\infty(\partial X^n, V(1^n_{\lambda-\zeta},{\varphi}))$$
be the restriction of sections. It is a $\Gamma$-equivariant and surjective
continuous linear map.
\begin{prop}\label{rrttee}
There exists a meromorphic family of maps
$$i^*_\Gamma:B_\Gamma(1^{n+1}_\lambda,{\varphi})\rightarrow
B_\Gamma(1^n_{\lambda-\zeta},{\varphi})$$
such that the following diagram commutes:
$$\begin{array}{ccc} C^\infty(\partial X,V(1^{n+1}_\lambda,{\varphi})) &\stackrel{i^*}{\rightarrow}&C^\infty(\partial X^n, V(1^n_{\lambda-\zeta},{\varphi}))\\
\downarrow \pi^{\Gamma,n+1}_*&&\downarrow \pi^{\Gamma,n}_*\\
B_\Gamma(1^{n+1}_\lambda,{\varphi})&\stackrel{i_\Gamma^*}{\rightarrow }& B_\Gamma(1^n_{\lambda-\zeta},{\varphi})
\end{array}\ .$$
If $\Gamma$ does not have cusps of full rank, then $i^*_\Gamma$ is holomorphic.
\end{prop}
This proposition will be proved in various stages. First we consider a version for the spaces associated to the cusps, see Lemma \ref{klopp}. The global result is stated in Lemma \ref{comppp}.
\subsubsection{}
In the following we explain the construction of $i_\Gamma^*$.
If all cusps of $\Gamma$ as a subgroup of $G^n$ have smaller rank, then
$i^*_\Gamma$ is induced
by the ususal restriction $i^*_\Gamma:
C^\infty(B_\Gamma,V_{B_\Gamma}(1_\lambda^{n+1},{\varphi}))\rightarrow C^\infty(B_\Gamma,V_{B_\Gamma}(1^n_{\lambda-\zeta},{\varphi}))$,
and both $\pi^{\Gamma,n}_*$
and $\pi^{\Gamma,n+1}_*$ are given as the average of sections
over $\Gamma$. The relation $i^*_\Gamma\circ \pi^{\Gamma,n+1}_*=\pi^{\Gamma,n}_*\circ i^*$
is obvious in the domain of convergence ${\rm Re }(\lambda)>\delta^{n+1}_\Gamma+\delta_{\varphi}$,
and follows by meromorphic continuation for all $\lambda\in{\aaaa_\C^\ast}$.
We postpone the proof of the fact that $i^*_\Gamma$
really maps $B_\Gamma(1^{n+1}_\lambda,{\varphi})$
to $B_\Gamma(1^{n}_{\lambda-\zeta},{\varphi})$
until these spaces are defined.
\subsubsection{}
If $\Gamma$ has cusps of full rank (when considered as a subgroup of $G^n$), then $i^*_\Gamma$
is the sum of two maps $(i^*_\Gamma)_1:B_\Gamma(1^{n+1}_\lambda,{\varphi})
\rightarrow B_\Gamma(1^n_{\lambda-\zeta},{\varphi})_1$ and $(i^*_\Gamma)_2:
B_\Gamma(1^{n+1}_\lambda,{\varphi})\rightarrow {R}_\Gamma(1^n_{\lambda-\zeta},{\varphi})_{max}$.
The map $(i^*_\Gamma)_1$ is the restriction of smooth sections, and
we can apply the argument above in order to show commutativity of
the part of the commutative
diagram involving $(\pi^{\Gamma,n}_*)_1$ and $(i^*_\Gamma)_1$.
\subsubsection{}
We define $(i^*_\Gamma)_2$ such that the corresponding diagram
is commutative. We choose $k\in{\mathbb {N}}_0$
so large that $E_{\infty_{P^n}}(1^n_{-\lambda+\zeta},\tilde{\varphi})
\subset C^{-k}(\partial X^n,V(1^n_{-\lambda+\zeta},\tilde{\varphi}))$
for all $P^n\in\tilde{\mathcal{P}}^{max,n}$. For $f\in B_\Gamma(1^{n+1}_\lambda,{\varphi})$
the component $i^*_\Gamma(f)_{2,P^n}
\in {R}^{U_{P^n}}(1^n_{\lambda-\zeta},{\varphi})$ is characterized
by $$\langle\phi,i^*_\Gamma(f)_{2,P^n}\rangle=\langle res^{U_{P^{n+1}},n+1}_k\circ i_*(\phi), T_{P^{n+1}}(f) \rangle$$
for all $\phi\in E_{\infty_{P^n}}(1^n_{-\lambda+\zeta},\tilde{\varphi})$,
where $P^{n+1}\subset G^{n+1}$ is the unique parabolic subgroup
containing $P^n$, and
$$T_{P^{n+1}}:C^\infty(B^{n+1}_\Gamma,V_{B_\Gamma}(1^{n+1}_\lambda,{\varphi}))\rightarrow C^\infty(B^{n+1}_{U_{P^{n+1}}},V_{B^{n+1}_{U_{P^{n+1}}}}(1^{n+1}_\lambda,{\varphi}))$$
is the natural map defined using the map $e_{P^{n+1}}$
and a cut-off function $\chi_p\in C^\infty(B_\Gamma)$
which is supported in $B_p$ and one on a neighbourhood of infinity
of $B_p$.
This definition is independent of the choice of $k$.
In order to show that $(i^*_\Gamma)_2\circ \pi^{\Gamma,n+1}_*=(\pi^{\Gamma,n}_*)_2\circ i^*$
we compute in the domain of convergence for generic $\lambda$ (so that the restrictions are defined)
\begin{eqnarray*}
\langle\phi,(i^*_\Gamma)_2\circ \pi^{\Gamma,n+1}_*(f)\rangle&=&
\langle res^{U_{P^{n+1}},n+1}_k\circ i_*(\phi),T^*_{P^{n+1}}\circ \pi^{\Gamma,n+1}_*(f)\rangle\\
&=&\sum_{g\in\Gamma^{P^{n+1}}} \langle res^{U_{P^{n+1}},n+1}_k\circ i_*(\phi),\chi_{P^{n+1}}\circ \pi^{U_{P^{n+1}},n+1}_*(\pi^{1^{n+1}_\lambda,{\varphi}}(g)f)\rangle\\
&\stackrel{(*)}{=}&\sum_{g\in\Gamma^{P^{n+1}}} \langle i_*(\phi), \pi^{1^{n+1}_\lambda,{\varphi}}(g)f\rangle\\
&=&\sum_{g\in\Gamma^{P^{n+1}}} \langle \pi^{1^{n+1}_{-\lambda+\zeta},\tilde{\varphi}}(g^{-1}) \phi, i^*(f)\rangle\\
&=&\langle \phi,(\pi^{\Gamma,n}_*)_2\circ i^*(f)\rangle\ ,
\end{eqnarray*}
where $\Gamma^{P^{n+1}}$ denotes any system of representatives
of $U_{P^{n+1}}\backslash \Gamma$, and $\chi_{P^{n+1}}:= e_{P^{n+1}}^* \chi_p$ (see Def.~\ref{t67} for the map $e_p$).
In order to get the equality marked by $(*)$ we use that
$$(1- \chi_{P^{n+1}})\circ res^{U_{P^{n+1}},n+1}_k \circ i_*(\phi)=0$$ and
$$ext^{U_{P^{n+1}},n+1}\circ res_k^{U_{P^{n+1}},n+1}\circ i_*(\phi)=i_*(\phi)\ .$$
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Let $i_*$ and $i_*^\Gamma$ denote the dual maps to
$i^*$ and $i^*_\Gamma$. Dualizing Prop.~\ref{rrttee} we obtain the following corollary.
\begin{kor}\label{rrttee1}
We have a commutative diagram of meromorphic families of maps
$$\begin{array}{ccc} C^{-\infty}(\partial X,V(1^{n}_\lambda,{\varphi})) &\stackrel{i_*}{\rightarrow}&C^{-\infty}(\partial X^{n+1}, V(1^{n+1}_{\lambda-\zeta},{\varphi}))\\
\uparrow ext^{\Gamma,n}_*&&\uparrow ext^{\Gamma,n+1}_*\\
D_\Gamma(1^{n}_\lambda,{\varphi})&\stackrel{i^\Gamma_*}{\rightarrow }&D_\Gamma(1^{n+1}_{\lambda-\zeta},{\varphi})
\end{array}\ .$$
\end{kor}
\subsubsection{}
\begin{lem}\label{jjjjjjjjjjjtwzew}
There exists holomorphic families of continuous maps
$$j^*:C^{-\infty}(\partial X^{n+1},V(1^{n+1}_{\lambda-\zeta},{\varphi}))\rightarrow C^{-\infty}(\partial X^n, V(1^n_\lambda,{\varphi}))$$
and $$j_*:
C^{\infty}(\partial X^n, V(1^n_{-\lambda},\tilde{\varphi}))\rightarrow
C^{\infty}(\partial X^{n+1},V(1^{n+1}_{-\lambda+\zeta},\tilde{\varphi}))$$
such that $j^*\circ i_* ={\mathrm {id}}$, $i^*\circ j_*={\mathrm {id}}$.
\end{lem}
{\it Proof.$\:\:\:\:$}
We choose a tubular neighbourhood $T:{\mathbb{F}}\times\partial X^n\hookrightarrow \partial X^{n+1}$
of $\partial X^n$ such that $T(\{0\}\times \partial X^n)=\partial X^n$.
Furthermore we choose a cut-off function $\chi\in C_c^\infty({\mathbb{F}})$
such that $\chi(0)=1$. Then we define the map
$t:
C^{\infty}(\partial X^n, V(1^n_{\rho^n},\tilde{\varphi}))\rightarrow
C^{\infty}(\partial X^{n+1},V(1^{n+1}_{\rho^{n+1}},\tilde{\varphi}))$
setting $(tf)(T(u,x)):=\chi(u)f(x)$ and extending
this function by zero outside the tubular neighbourhood.
Here we use the canonical identifications
$C^{\infty}(\partial X^n, V(1^n_{\rho^n},\tilde{\varphi}))=C^{\infty}(\partial X^n)\otimes V_{\tilde{\varphi}}$
and $C^{\infty}(\partial X^{n+1},V(1^{n+1}_{\rho^{n+1}},{\varphi}))=C^{\infty}(\partial X^{n+1})\otimes V_{\tilde{\varphi}}$.
We choose a positive section
$s_{n+1}\in C^{\infty}(\partial X^{n+1},V(1^{n+1}_{\rho^{n+1}+\alpha}))$ and put $i^* s_{n+1}=:s_n\in C^{\infty}(\partial X^n, V(1^n_{\rho^n+\alpha}))$.
Then we define
$\Phi_{\rho^n}:C^{\infty}(\partial X^n, V(1^n_{-\lambda},\tilde{\varphi}))
\rightarrow C^{\infty}(\partial X^n, V(1^n_{\rho^n},\tilde{\varphi}))$
and
$\Phi_{\rho^{n+1}}:
C^{\infty}(\partial X^{n+1},V(1^{n+1}_{-\lambda+\zeta},\tilde{\varphi}))\rightarrow
C^{\infty}(\partial X^{n+1},V(1^{n+1}_{\rho^{n+1}},\tilde{\varphi}))$
as multiplication by $s_n^{(\rho^n+\lambda)/\alpha}$ and
$s_{n+1}^{(\rho^{n+1}+\lambda-\zeta)/\alpha}$, respectively.
We define $j_*:=\Phi_{\rho^{n+1}}^{-1}\circ t\circ \Phi_{\rho^n}$,
and $j^*$ as the adjoint of $j_*$. It is now easy to check
that $j_*$ and $j^*$ have the required properties. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}\label{ohu}
If $G^n$ does not belong to the list
$\{Spin(1,n),SO(1,n)_0,SU(1,n),Sp(1,n)\}$, then there is a finite covering
$p:\tilde G^n\rightarrow G^n$ with $\tilde G^n\in \{Spin(1,n),SO(1,n)_0,SU(1,n),Sp(1,n)\}$.
In this case can find a normal subgroup
$\Gamma^0\subset \Gamma$ of finite index and a discrete subgroup $\tilde \Gamma^0\subset \tilde G^n$ such that $p$ induces an isomorphism from $\tilde \Gamma^0$ to $\Gamma^0$. Indeed, using Selberg's Lemma we can take
a torsion-free subgroup $\tilde\Gamma^0$ of $p^{-1}(\Gamma)$ of finite index
and set $\Gamma^0:=p(\tilde\Gamma^0)$.
We can apply the concept of embedding to
the subgroup $\tilde\Gamma^0$. In order to transfer results for
$\tilde\Gamma^0$ to $\Gamma$ we use averages over the finite group $\Gamma/\Gamma^0$.
\subsection{Twisting}\label{ttw}
\subsubsection{}
Twisting is an important technical device of the present paper. We explain in \ref{weih125} and \ref{weih1266} how this concept is used. But first we must introduce some notation.
\subsubsection{}\label{weih122}
If $(\pi,V_\pi)$ is a finite-dimensional representation
of $G$, then we can form
the bundles $V(\sigma_\lambda\otimes\pi,{\varphi})$ and
$V(\sigma_\lambda,\pi\otimes {\varphi})$.
There is an isomorphism $T:V(\sigma_\lambda\otimes\pi,{\varphi})\stackrel{\sim}{\rightarrow}
V(\sigma_\lambda,\pi\otimes {\varphi})$, which is given on the level of sections by
$T:C^\infty(\partial X,V(\sigma_\lambda\otimes\pi,{\varphi}))
\stackrel{\sim}{\rightarrow} C^\infty(\partial X,V(\sigma_\lambda,\pi\otimes{\varphi}))$, $T(f)(g):=\pi(g)f(g)$.
\subsubsection{}\label{weih144}
Given an irreducible representation $\sigma$ of $M$
and $\mu_0\in{\mathfrak{a}}^*$ there exists a finite-dimensional representation $(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$
of $G$ with highest $A$-weight $\mu\ge\mu_0$ such that $V_\sigma=V_{\pi_{\sigma,\mu}}(\mu)$
as representations of $M$. Here $V_{\pi_{\sigma,\mu}}(\mu)$ denotes the subspace
of $V_{\pi_{\sigma,\mu}}$ on which $A$ acts with weight $\mu$.
Note that $V_{\pi_{\sigma,\mu}}(-\mu)= V_{\sigma^w}$
as representations of $M$.
\subsubsection{}\label{weih241}
If $\sigma$ is a representation of $M$ of the form
$\sigma^\prime\oplus (\sigma^\prime)^w$, where $\sigma^\prime$
irreducible and not equivalent to its Weyl conjugate, then given $\mu_0\in{\mathfrak{a}}^*$
there exist
$(\pi_{\sigma^\prime,\mu},V_{\pi_{\sigma^\prime,\mu}})$ and $(\pi_{(\sigma^\prime)^w,\mu},V_{\pi_{(\sigma^\prime)^w,\mu}})$
for suitable $\mu\ge \mu_0$. In this case we set
$\pi_{\sigma,\mu}:=\pi_{\sigma^\prime,\mu} \oplus \pi_{(\sigma^\prime)^w,\mu}$.
In connection with twisting without further notice we will always assume that
$\sigma$ is Weyl-invariant and either irreducible or of the form
$\sigma^\prime\oplus (\sigma^\prime)^w$, where $\sigma^\prime$
is irreducible and not Weyl invariant.
Note that $\pi_{\tilde\sigma,\mu}=\tilde\pi_{\sigma,\mu}$.
\subsubsection{}\label{weih142}
We have an embedding of $P$-modules
$V_{\sigma_\lambda}\hookrightarrow V_{1_{\lambda+\mu}}\otimes V_{\pi_{\sigma,\mu}}$
and a corresponding embedding of $\Gamma$-equivariant bundles
$V(\sigma_\lambda,{\varphi})\hookrightarrow V(1_{\lambda+\mu}\otimes\pi_{\sigma,\mu},{\varphi})$.
Composing this embedding with the isomorphism $T$ (introduced in \ref{weih122})
we obtain the $\Gamma$-equivariant embedding
$$i_{\sigma,\mu}:C^{\pm\infty}(\partial X,V(\sigma_\lambda,{\varphi}))\hookrightarrow
C^{\pm\infty}(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\ .$$
Similarly we have a projection of $P$-modules
$V_{1_{\lambda-\mu}}\otimes V_{\pi_{\sigma,\mu}} \rightarrow V_{\sigma_{\lambda}}$ and a corresponding projection of $\Gamma$-equivariant bundles
$V(1_{\lambda-\mu}\otimes\pi_{\sigma,\mu},{\varphi})\rightarrow V(\sigma_{\lambda},{\varphi})$. Composing this projection with the inverse of $T$
we obtain the $\Gamma$-equivariant projection
$$p_{\sigma,\mu}:C^{\pm\infty}(\partial X,V(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi}))
\rightarrow C^{\pm\infty}(\partial X,V(\sigma_{\lambda},{\varphi}))\ .$$
\subsubsection{}\label{neuj1000}
Since $i_{\sigma,\mu}$ and $p_{\sigma,\mu}$ are induced by homomorphisms
of $\Gamma$-equivariant bundles restriction to $\Gamma$-equivariant
sections over $\Omega_\Gamma$ provides
the embedding
$$i^\Gamma_{\sigma,\mu}:C^{\infty}(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))\hookrightarrow
C^{\infty}(B_\Gamma,V_{B_\Gamma}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))$$
and the projection
$$p^\Gamma_{\sigma,\mu}:C^{\infty}(B_\Gamma,V_{B_\Gamma}(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi}))
\rightarrow C^{\infty}(B_\Gamma,V_{B_\Gamma}(\sigma_{\lambda},{\varphi}))\ .$$
In Lemma \ref{compat3} will show that $i^\Gamma_{\sigma,\mu}$ (resp. $p^\Gamma_{\sigma,\mu}$)
maps $B_\Gamma(\sigma_{\lambda},{\varphi})_1$ (resp. $B_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi})_1$)
to $B_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})_1$
(resp. $B_\Gamma(\sigma_{\lambda},{\varphi})_1$).
Hence we can consider the
adjoint of
$$p^\Gamma_{\tilde\sigma,\mu}:B_\Gamma(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi})_1\rightarrow B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})_1$$ which will be denoted by
by $$(i^\Gamma_{\sigma,\mu})_1:D_\Gamma(\sigma_{\lambda},{\varphi})_1
\rightarrow D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})_1\ .$$
In a similar manner the adjoint of
$$i^\Gamma_{\tilde\sigma,\mu}:B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})_1\rightarrow B_\Gamma(1_{-\lambda+\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi})_1$$
will be a map
$$(p^\Gamma_{\sigma,\mu})_1:D_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi})_1\rightarrow D_\Gamma(\sigma_{\lambda},{\varphi})_1\ .$$
\subsubsection{}
In the presence of cusps of full rank we must take into account their contribution to the function spaces (see \ref{weih123}). It is clear that $i_{\sigma,\mu}$ (resp. $p_{\sigma,\mu}$) maps
$E_{\infty_P}(\sigma_\lambda,{\varphi})$ (resp. $E_{\infty_P}(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi})$) to $E_{\infty_P}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$ (resp.
$E_{\infty_P}(\sigma_{\lambda},{\varphi})$).
We get second components
\begin{eqnarray*}
(i^\Gamma_{\sigma,\mu})_2&:&{R}_{\Gamma}(\sigma_\lambda,{\varphi})_{max}\to {R}_{\Gamma}(1_{\lambda+\mu},\pi_{\sigma,\mu}{\varphi})_{max}\\(p^\Gamma_{\sigma,\mu})_2&:&{R}_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi})_{max} \rightarrow {R}_\Gamma(\sigma_{\lambda},{\varphi})_{max}\end{eqnarray*}
Combining all these maps we obtain
maps
\begin{eqnarray*}
i^\Gamma_{\sigma,\mu}&:&D_\Gamma(\sigma_{\lambda},{\varphi})\rightarrow D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})\\
p^\Gamma_{\sigma,\mu}&:&D_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi})\rightarrow D_\Gamma(\sigma_{\lambda},{\varphi})
\end{eqnarray*}
and dually corresponding maps between spaces
of functions.
\subsubsection{}
Using the isomorphism $T$ we transfer the action
$\pi^{1_{-\lambda-\mu}\otimes\pi_{\tilde\sigma,\mu},{\mathrm {id}}_{\tilde{\varphi}}}$ of ${\mathcal{Z}}({\mathfrak{g}})$ to $$C^\infty(\partial X,V(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes \tilde{\varphi}))\ .$$ Since this action commutes with $\Gamma$ and
is implemented by local operators we obtain
an action of ${\mathcal{Z}}({\mathfrak{g}})$ on
$C^\infty(B_\Gamma,V_{B_{\Gamma}}(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes \tilde{\varphi}))$ as well.
We will show that $B_\Gamma(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi})_1$
is a $Z({\mathfrak{g}})$-invariant subspace of $C^\infty(B_\Gamma,V_{B_{\Gamma}}(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes \tilde{\varphi}))$. By duality we obtain an action of $Z({\mathfrak{g}})$
on $D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})_1$.
Since ${\mathcal{Z}}({\mathfrak{g}})$ also acts on
$E_{\infty_P}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$, $P\in \tilde{\mathcal{P}}^{max}$, the space
$D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$
has the structure of a ${\mathcal{Z}}({\mathfrak{g}})$-module.
\subsubsection{}\label{weih146}
Let $\Omega\in{\mathcal{Z}}({\mathfrak{g}})$ denote the Casimir operator.
Let $\{\sigma^i | i\in I_\nu\}$, be the set of Weyl-invariant
representations of $M$ (irreducible or sum of two non-Weyl invariant
irreducible) occuring in the restriction
of $V_{\pi_{\sigma,\mu}}(\nu)$ to $M$, where we set $I_\mu:=\{1\}$.
Let
$\nu_i(\lambda):=\pi^{\sigma^i_{\lambda+\nu}}(\Omega)\in{\mathbb {C}}$
be the eigenvalue of the Casimir operator on the principal series representation
$\pi^{\sigma^i_{\lambda+\nu}}$. We set $I:=\bigcup_{\nu\not=\mu} I_\nu$
and define the meromorphic function
$${\aaaa_\C^\ast}\ni\lambda\mapsto \Pi(\lambda):=1-\prod_{i\in I} \frac{\Omega-\nu_i(\lambda)}{\nu_1(\lambda)-\nu_i(\lambda)}\in{\mathcal{Z}}({\mathfrak{g}})\ .$$
This function has a finite number of poles all contained in the subset
$I_{\mathfrak{a}}\subset {\mathfrak{a}}^*$.
Furthermore, we define
$$Z(\lambda):=\pi^{1_{\lambda+\mu}\otimes\pi_{\sigma,\mu},{\mathrm {id}}_{{\varphi}}}(\Pi(\lambda))\ .$$
If $\lambda\not\in I_{\mathfrak{a}}$ (or more precisely, if $\Pi(\lambda)$ is regular),
then $Z(\lambda)$ is a projection.
\subsubsection{}
For $\lambda\in I_{\mathfrak{a}}$ we will show in Lemma \ref{compat6}
that
$$i_{\sigma,\mu}^\Gamma:D_\Gamma(\sigma_{\lambda},{\varphi})\rightarrow
\ker \{Z(\lambda):D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})\rightarrow
D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})\}=:\ker_\Gamma(Z(\lambda))$$
is an isomorphism. We then define the continuous map
$$j_{\sigma,\mu}^\Gamma:D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})
\stackrel{1-Z(\lambda)}{\longrightarrow} \ker_\Gamma(Z(\lambda))\stackrel{(i_{\sigma,\mu}^\Gamma)^{-1}}{\longrightarrow}
D_\Gamma(\sigma_{\lambda},{\varphi})\ .$$
In Lemma \ref{compat6} we will furthermore show that as a function of $\lambda$ the maps
$j_{\sigma,\mu}^\Gamma$ form a meromorphic family of continuous maps
which is holomorphic on ${\aaaa_\C^\ast}\setminus I_{\mathfrak{a}}$.
If $\Gamma$ is trivial, then we omit the superscript
and write $j_{\sigma,\mu}$ for $j_{\sigma,\mu}^{\{1\}}$.
\subsubsection{}\label{weih125}
We now explain by examples how the concept of twisting is applied.
Assume that we have obtained a meromorphic continuation of the extension
$ext^\Gamma:D_\Gamma(1_\lambda,{\varphi})\rightarrow C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$ to some half-plane
$W=\{{\rm Re }(\lambda)>\lambda_0\}$ for all admissible twists ${\varphi}$.
Then we can obtain a meromorphic continuation of
$ext^\Gamma:D_\Gamma(\sigma_\lambda,{\varphi})\rightarrow
C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$ to some larger half plane
$W-\mu_0$, $\mu_0\ge 0$, and for all $\sigma$ as follows. We choose $(\pi_{\sigma,\mu}, V_{\pi_{\sigma,\mu}})$ for some $\mu\ge \mu_0$.
Then we can define a meromorphic family
$\widetilde{ext}^\Gamma:D_\Gamma(\sigma_\lambda,{\varphi})\rightarrow
C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$
for $\lambda\in W-\mu_0$ by the diagram
$$\begin{array}{ccc}
D_\Gamma(\sigma_\lambda,{\varphi})&\stackrel{i^\Gamma_{\sigma,\mu}}{\rightarrow}&D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\\
\downarrow \widetilde{ext}^\Gamma&&\downarrow ext^\Gamma\\
C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))&\stackrel{j_{\sigma,\mu}}{\leftarrow}&
C^{-\infty}(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))
\end{array}\ .$$
We then show that $ext^\Gamma=\widetilde{ext}^\Gamma$
for all $\lambda$ with sufficiently large real part.
Thus $\widetilde{ext}^\Gamma$ provides a meromorphic continuation
of $ext^\Gamma$.
\subsubsection{}\label{weih1266}
Assume that we have defined the scattering matrix
$S^\Gamma_\lambda:D_\Gamma(1_\lambda,{\varphi})\rightarrow D_\Gamma(1_{-\lambda},{\varphi})$
in the spherical case for all admissible twists ${\varphi}$.
Then we could define the scattering matrix
$S^\Gamma_\lambda:D_\Gamma(\sigma_\lambda,{\varphi})\rightarrow D_\Gamma(\sigma_{-\lambda},{\varphi})$ for all $\sigma$ using the diagram
\begin{equation}\label{qw11}
\begin{array}{ccc}
D_\Gamma(\sigma_\lambda,{\varphi})&\stackrel{i^\Gamma_{\sigma,\mu}}{\rightarrow}&D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})\\
\downarrow S^\Gamma_\lambda&&\downarrow S^\Gamma_{\lambda+\mu}\\
D_\Gamma(\sigma_{-\lambda},{\varphi})&\stackrel{p^\Gamma_{\sigma,\mu}}{\leftarrow}&D_\Gamma(1_{-\lambda-\mu},\pi_{\sigma,\mu}\otimes{\varphi})
\end{array}\ .
\end{equation}
For generic $\lambda\in{\aaaa_\C^\ast}$ we have a well-defined map
$res^\Gamma:{}^\Gamma C^{-\infty}(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))
\rightarrow D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$,
and we can define
$res^\Gamma:{}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))
\rightarrow D_\Gamma(\sigma_\lambda,{\varphi})$ using the diagram
$$\begin{array}{ccc}
{}^\Gamma C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi})) &\stackrel{i_{\sigma,\mu}}{\rightarrow}&{}^\Gamma C^{-\infty}(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\\
\downarrow res^\Gamma &&\downarrow res^\Gamma \\
D_\Gamma(\sigma_{\lambda},{\varphi})&\stackrel{j^\Gamma_{\sigma,\mu}}{\leftarrow}&D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})
\end{array}\ .$$
We then check that
$S^\Gamma_\lambda$ defined by (\ref{qw11}) coincides with
$res^\Gamma\circ J_\lambda\circ ext^\Gamma$.
\subsection{Holomorphic vector bundles over ${\mathbb {C}}$}
\subsubsection{}
It is a general fact
that each finite-dimensional
holomorphic vector bundle over ${\mathbb {C}}$ or a half-plane is trivial.
Let $U\subset {\mathbb {C}}$ be open and $E$ be a finite-dimensional holomorphic vector bundle
over $U$. Furthermore let $U\times V\rightarrow U$ be a trivial
bundle of Fr\'echet spaces over $U$. By ${\mathcal{E}}$ and ${\cal V}$ we denote
the corresponding sheaves of holomorphic sections.
Let $\phi:E\rightarrow V$ be a meromorphic family of linear maps.
The assertions of the following lemma are well-known in the case where $V$ is finite-dimensional.
\begin{lem}\label{bunle}
\begin{enumerate}
\item
There exists a unique finite-dimensional subbundle $F\subset U\times V$
such that $\phi$ factors over a meromorphic family of maps $\psi:E\rightarrow F$
which is surjective for generic $z\in U$.
\item
If we have meromorphic families of bundle maps
$Z_E:E\rightarrow E$ and $Z_V:V\rightarrow V$ such that $\phi\circ Z_E=Z_V\circ \phi$, then $Z_V$ restricts to a meromorphic family of
bundle maps of $F$.
\item
Furthermore, if $U={\mathbb {C}}$ or a half-plane, then there exists a meromorphic right-inverse
$\eta:F\rightarrow E$ such that $\phi\circ \eta={\mathrm {id}}_F$.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
The proof of this lemma will occupy the remainder of the present subsection.
The main point which we will explain in detail is the reduction to the finite-dimensional case.
\subsubsection{}
We first show that $F$ exists.
Let $\phi$ have singularities in the discrete subset
$A\subset U$, and let $n_z$, $z\in A$,
denote the order of the corresponding singularity.
We consider the divisor $D:=\sum_{z\in A} n_z z$, the associated line bundle
$L(D)$, and its sheaf of sections ${\cal L}(D)$.
Let ${\cal V}(D)$ be the sheaf of sections of $V\otimes L(D)$.
Then $\phi$ induces a holomorphic map
$\phi_D:E\rightarrow V\otimes L(D)$. Let ${\mathcal{ W}}$ be the quotient sheaf
$${\mathcal{E}} \stackrel{\phi_D}{\rightarrow}{\cal V}(D)\rightarrow {\mathcal{ W}}\rightarrow 0$$
and denote by $Tor({\mathcal{ W}})$ its torsion subsheaf. We define
${\mathcal{F}}(D)$ as the kernel
$$0\rightarrow {\mathcal{F}}(D)\rightarrow {\cal V}(D)\rightarrow {\mathcal{ W}}/Tor({\mathcal{ W}})\rightarrow 0\ .$$
There is a natural factorization of $\phi_D$ over $\psi_D:{\mathcal{E}}\rightarrow {\mathcal{F}}(D)$.
\subsubsection{}\label{weih129}
We claim that ${\mathcal{F}}(D)$ is a coherent sheaf.
Let $x\in U$.
If $v\in {\cal V}(D)_x$, then let $lp(v)\in V$ denote
the leading part of $v$ at $x$. Note that $lp(v)=0$ implies that $v=0$ since a Laurent series must start somewhere.
We define the subspace
$Z\subset V$ as the set of all leading parts $lp(\phi_D(e))$, $e\in {\mathcal{E}}_x$.
\subsubsection{}
We now show that $\dim(Z)<\infty$.
We choose a connected neighbourhood $U_x\subset U$
of $x$ and a holomorphic trivialization
$E_{|U_x}=U_x\times E_x$. Let $\phi=\sum_{n\ge m} \phi_n z^n$
denote the Laurent expansion of $\phi$ in this trivialization, where
$\phi_n\in {\mbox{\rm Hom}}(E_x,V)$. Then using that $\dim E_x<\infty$ we have
$$Z=\phi_m(E_x)+\phi_{m+1}(\ker\phi_m)+\phi_{m+2}(\ker\phi_m\cap\ker\phi_{m+1})+\dots + \phi_{m+k}(\bigcap_{i=0}^{k-1} \ker\phi_{m+i})\ ,$$
where $k\in{\mathbb {N}}_0$ is sufficiently large such that
$\bigcap_{i=0}^{k-1} \ker\phi_{m+i}=\bigcap_{i=0}^{l-1} \ker\phi_{m+i}$
for all $l\ge k$. This proves that $\dim(Z)<\infty$.
\subsubsection{}
We choose some closed subspace $Y\subset V$ of finite codimension
such that $Z\cap Y=\{0\}$.
We consider the composition
$\overline{\phi}_{D}:E\stackrel{\phi}\rightarrow V \rightarrow V/Y$.
We form the quotient of sheaves
$${\mathcal{E}} \stackrel{\overline{\phi}_D}{\rightarrow}{\cal V}(D)/{\cal Y}(D)\rightarrow\tilde {\mathcal{ W}}\rightarrow 0\ ,$$
and we define $\tilde {\mathcal{F}}(D)$ as the kernel
$$0\rightarrow\tilde{\mathcal{F}}(D)\rightarrow {\cal V}(D)/{\cal Y}(D)\rightarrow\tilde {\mathcal{ W}}/Tor(\tilde {\mathcal{ W}})\rightarrow 0\ .$$
\subsubsection{}
Consider the following commutative diagram of sheaves on $U_x$:
$$\begin{array}{ccccccccc}
&&0&&0&&0&&\\
&&\downarrow&&\downarrow&&\downarrow&&\\
0&\rightarrow&0&\rightarrow&{\cal Y}(D)&\stackrel{\cong}{\rightarrow}& {\cal Y}(D)&\rightarrow &0\\
&&\downarrow&&\downarrow&&\downarrow&&\\
0&\rightarrow&{\mathcal{E}}/\ker \phi_D&\rightarrow&{\cal V}(D)&\rightarrow& {\mathcal{ W}}&\rightarrow &0\\
&&\downarrow\cong &&\downarrow&&\downarrow&&\\
0&\rightarrow&{\mathcal{E}}/\ker \overline{\phi}_D&\rightarrow&{\cal V}(D)/{\cal Y}(D)&\rightarrow& \tilde{\mathcal{ W}}&\rightarrow &0\\
&&\downarrow&&\downarrow&&\downarrow&&\\
&&0&&0&&0&&\end{array}
$$
The two lower rows are exact by construction.
We now show that the left column is exact.
We have a sequence
$$\ker(\phi_D)\subset \ker(\overline{\phi_D})\subset {\mathcal{E}}$$ of inclusioins of torsion-free sheaves.
In particular, $\ker(\phi_D)$ and $\ker(\overline{\phi_D})$ are sheaves of holomorphic sections of vector bundles on $U_x$. If we show that the inclusion induces an isomorphism
\begin{equation}\label{weih128}
\ker(\phi_D)_x\cong \ker(\overline{\phi_D})_x\ ,
\end{equation}
then we conclude an isomorphism of sheaves
$\ker(\phi_D) \cong \ker(\overline{\phi_D})$ after shrinking $U_x$, if necessary.
To see (\ref{weih128}) consider $h\in \ker (\overline{\phi}_D)_x$. Then $lp(\phi_D(h))\in Y$.
Since on the other hand $lp(\phi_D(h))\in Z$ we conclude that
$lp(\phi_D(h))=0$, hence $h\in\ker (\phi_D)_x$.
We now conclude that
the last column is exact, too.
\subsubsection{}
We now consider the diagram
$$\begin{array}{ccccccccc}
&&0&&0&&0&&\\
&&\downarrow&&\downarrow&&\downarrow&&\\
0&\rightarrow&0&\rightarrow&Tor {\mathcal{ W}}&\stackrel{\cong}{\rightarrow}& Tor\tilde {\mathcal{ W}} &\rightarrow &0\\
&&\downarrow&&\downarrow&&\downarrow&&\\
0&\rightarrow&{\cal Y}(D)&\rightarrow&{\mathcal{ W}}&\rightarrow&\tilde {\mathcal{ W}}&\rightarrow &0\\
&&\downarrow\cong &&\downarrow&&\downarrow&&\\
0&\rightarrow&{\cal Y}(D)&\rightarrow&{\mathcal{ W}} /Tor {\mathcal{ W}}&\rightarrow&\tilde {\mathcal{ W}}/ Tor\tilde{\mathcal{ W}}&\rightarrow &0\\
&&\downarrow&&\downarrow&&\downarrow&&\\
&&0&&0&&0&&\end{array}\ .
$$
Since ${\cal Y}(D)$ is torsion-free we have ${\cal Y}(D)\cap Tor{\mathcal{ W}}
=0$ This implies the isomorphism in the upper row.
It follows that all rows and columns of this diagram are exact.
\subsubsection{}
We now consider the following diagram of sheaves on $U_x$:
$$\begin{array}{ccccccccc}
&&0&&0&&0&&\\
&&\downarrow&&\downarrow&&\downarrow&&\\
0&\rightarrow&0&\rightarrow&{\cal Y}(D)&\stackrel{\cong}{\rightarrow}& {\cal Y}(D)&\rightarrow &0\\
&&\downarrow&&\downarrow&&\downarrow&&\\
0&\rightarrow&{\mathcal{F}}(D)&\rightarrow&{\cal V}(D)&\rightarrow& {\mathcal{ W}}/Tor {\mathcal{ W}}&\rightarrow &0\\
&&\downarrow\cong &&\downarrow&&\downarrow&&\\
0&\rightarrow&\tilde{\mathcal{F}}(D)&\rightarrow&{\cal V}(D)/{\cal Y}(D)&\rightarrow& \tilde{\mathcal{ W}}/Tor \tilde{\mathcal{ W}}&\rightarrow &0\\
&&\downarrow&&\downarrow&&\downarrow&&\\
&&0&&0&&0&&\end{array}
$$
The rows and the middle column are exact. We just have shown
that the right column is exact.
\subsubsection{}
We conclude that the natural map ${\mathcal{F}}(D)\rightarrow \tilde {\mathcal{F}}(D)$
is an isomorphism of sheaves.
This proves the claim \ref{weih129} since
$\tilde {\mathcal{F}}(D)$ is obviously coherent.
\subsubsection{}
Since ${\mathcal{F}}(D)$ is torsion-free it is the sheaf of sections
of a holomorphic vector bundle $F(D)$ (here we use the fact that
the base space is smooth and one-dimensional).
By construction ${\mathcal{F}}(D)/\psi_D({\mathcal{E}})$ is torsion and therefore
$\psi_{D,x}:E_x\rightarrow F(D)_x$ is surjective for
generic $x\in U$.
We define $F:=F(D)\otimes L(-D)$ and let $\psi:E\rightarrow F$
be the corresponding meromorphic family of maps.
The uniqueness part and assertion 2.
are left to the reader.
\subsubsection{}
We now construct the meromorphic right-inverse $\eta$.
Note that we can assume that $E$ and $F$ are trivial.
We fix trivializations and a constant Hermitian metric
on $E$. Let $z\in {\mathbb {C}}$ be such that $\phi:E_z\rightarrow F_z$
is regular and surjective. Let $P$ be the orthogonal projection onto the
orthogonal complement of $\ker\phi_z$.
We extend $P$ constantly over $U$ and let $P(E)$ be the
range of $P$.
The composition $\phi\circ P:P(E)\rightarrow F$ is invertible at $z$
and therefore has a meromorphic family of inverses
$\eta:F\rightarrow P(E)\subset E$.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
\begin{ddd}\label{imagebundle}
The bundle $F$ constructed in Lemma \ref{bunle} is called the
image bundle of $\phi$.
\end{ddd}
\section{Pure cusps}
\subsection{Geometry of cusps}\label{cuspgeom}
\subsubsection{}
In this subsection we analyze the geometry of cusps.
First we recall the following theorem of Auslander.
\begin{theorem}[\cite{auslander61}, \cite{MR1490024}] \label{ausl}
Let $N$ be a connected, simply-connected nilpotent Lie group and $M$
be a compact group of automorphisms of $N$.
Furthermore let $U\subset N\rtimes M$ be a discrete subgroup
and $U^*:=\overline{NU}_0\cap U=p_{|U}^{-1}(\overline{p(U)}_0)$,
where $p:N\rtimes M\rightarrow M$ is the projection,
$()_0$ stands for connected component of the identity,
and "$\bar{.}$" means the closure of the set in the argument.
Then
\begin{enumerate}
\item $U^*\subset U$ is a normal subgroup of finite index.
\item There exists $b\in N$ and a connected subgroup
$N_V\subset N$ such that $(U^*)^b:=bU^*b^{-1}$ acts
effectively and cocompactly by translations on
$N_V$, i.e. there is lattice $V^*\subset N_V$ and an isomorphism
$\theta:(U^*)^b\rightarrow V^*$ such that
$u.x = \theta(u)x$, $x\in N_V$, where $(N\rtimes M)\times N\ni (a,x)\mapsto a.x\in N$
denotes the natural action of $N\rtimes M$ on $N$.
\item The element $b\in N$ of 2. can be choosen
such that $U^b$ leaves the space $N_V$ invariant.
\item $M_{U^0}:=\overline{p(U^*)}\subset M$ is a torus.
\end{enumerate}
\end{theorem}
The third assertion is due to Apanasov \cite{MR1490024}.
\subsubsection{}
We now apply Theorem \ref{ausl}
to a discrete torsion-free subgroup $U\subset P$
of a parabolic subgroup $P\subset G$ such that
$P$ is $U$-cuspidal. We have the exact sequence
\begin{equation}\label{yy77}
0\rightarrow N\rightarrow P\stackrel{l}{\rightarrow} L\rightarrow 0\ ,\end{equation}
where $N$ is the nil-radical of $P$. The group
$L$ decomposes as $L=M\times A$, and we denote
by $l_M:P\rightarrow M$ the composition of
$l$ with the projection from $L$ to $M$.
Let $E:=l^{-1}(M\times\{1\})$.
If we choose a split $s_1:M\rightarrow E$ of the sequence
$0\rightarrow N\rightarrow E\stackrel{l_M}{\rightarrow} M\rightarrow 0$,
then $m\in M$ acts by the automorphism $s_1(m)$ on $N$, and
we have an isomorphism $T_{s_1}:E\rightarrow N\rtimes_{s_1}M$.
Since $P$ is $U$-cuspidal we have $U\subset E$.
Applying Theorem \ref{ausl} to $T_{s_1}(U)$ we obtain a subgroup
$T_{s_1}(U)^*\subset T_{s_1}(U)$ of finite index, $b\in N$, a connected
subgroup $N_V$, a lattice $V^*\subset N_V$, and an isomorphism
$\theta:bT_{s_1}(U)^*b^{-1}\rightarrow V^*$
such that $bT_{s_1}(U)b^{-1}$ leaves $N_V$ invariant and acts by translations on $N_V$ via $\theta$.
\subsubsection{}
If we replace the split $s_1$ and $N_V$ by $b^{-1}N_Vb$ and
$s:=b^{-1}s_1b$, then $T_s(U)$ itself leaves $N_V$
invariant, and $T_s(U)^*$ acts by translations on $N_V$.
From now on we use the split $s$ in order to identify $M$ with the
subgroup $s(M)\subset E\subset P$, to write $E$ as the product $NM$,
and to indentify $T_s(U)$ with $U$.
\subsubsection{}\label{weih131}
Any element $u\in U^*$ ($u\in U$) can be written as $n_um_u\in NM$
such that $n_u\in N_V$ and $m_u$ centralizes (normalizes) $N_V$.
We have $V^*:=\{n_u|u\in U^*\}\subset N_V$.
The map $V^*\ni n_u \mapsto m_u\in M$ defines a homomorphism
$m^*:V^*\rightarrow M_{U^0}$.
\begin{lem}
There exists a subgroup $V\subset V^*$ of finite index
such that the restriction $m:=m^*_{|V}$ extends to a homomorphism
$m:N_V\rightarrow M_{U^0}$.
\end{lem}
{\it Proof.$\:\:\:\:$}
Let $\pi:V^*\rightarrow N_V/[N_V,N_V]$ be the projection.
Then $\pi(V^*)\subset N_V/[N_V,N_V]$ is a lattice (see
e.g. \cite{raghunathan72}, proof of Thm. 2.10).
Let $g_1,\dots g_n\in V^*$, $n=\dim(N_V/[N_V,N_V])$
be such that $\pi(g_1),\dots,\pi(g_n)$ generate the lattice
$\pi(V^*)$. We define $V:=\langle g_1,\dots,g_n\rangle$.
Furthermore we put $X_i:=\log(g_i)\in{\mathfrak{n}}_V$
and consider $\pi(X_i)\in {\mathfrak{n}}_V/[{\mathfrak{n}}_V,{\mathfrak{n}}_V]$.
Then $\pi(X_i)$ is a basis of ${\mathfrak{n}}_V/[{\mathfrak{n}}_V,{\mathfrak{n}}_V]$,
and $\{X_1,\dots,X_n\}$ generate the Lie algebra ${\mathfrak{n}}_V$.
We conclude that $N_V$ is the smallest connected group
containing $V$, and by \cite{raghunathan72}, Ch.2, $V$ is a lattice
in $N_V$. Therefore $V\subset V^*$ has finite index.
We claim that $V\cap [N_V,N_V]=[V,V]$.
Let $g\in V\cap [N_V,N_V]$.
Then there are finite sequences $i_k\in\{1,\dots,n\}$ and $e_k\in\{1,-1\}$, $k=1,\dots r$,
such that $g=g_{i_1}^{e_1}\dots g_{i_r}^{e_r}$.
Applying $\pi$ we obtain
$1=\pi(g_{i_1})^{e_1}\dots \pi(g_{i_r})^{e_r}$.
Using the fact that $N_V/[N_V,N_V]$ is free abelian we conclude
that there is a permutation $\sigma\in S_r$ such that
$1=g_{i_{\sigma(1)}}^{e_{\sigma(1)}}\dots g_{i_{\sigma(r)}}^{e_{\sigma(r)}}$.
We conclude that
$1=g$ modulo $[V,V]$, i.e. $g\in [V,V]$. This proves the claim.
We first define the derivative $dm:{\mathfrak{n}}_V\rightarrow {\mathfrak{m}}_{U^0}$
of $m$ as the composition
$${\mathfrak{n}}_V\stackrel{\pi}{\rightarrow} {\mathfrak{n}}_V/[{\mathfrak{n}}_V,{\mathfrak{n}}_V]\stackrel{q}{\rightarrow} {\mathfrak{m}}_{U^0}\ ,$$
where $q$ is given by $q(\pi(X_i)):= Z_i$, $i=1,\dots n$,
where $Z_i\in{\mathfrak{m}}_{N_V}$ is any element satisfying $\exp(Z_i)=m(g_i)$.
Integrating the derivative we obtain
a representation $\tilde m:N_V\rightarrow M_{U^0}$
such that $m(g_i)=\tilde m(g_i)$, $i=1\dots,n$.
Since $M_{U^0}$ is a torus we see that
$m$ must factor over $[V,V]$ and conclude that $m=\tilde m$.
This finishes the proof of the lemma.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}\label{weih251}
We consider the subgroup of finite index $U^1:=\{v m(v)| v\in V\}\subset U$.
\begin{ddd}\label{tu0}
We define $U^0$ to be the largest normal subgroup of $U$
such that $U^0\subset U^1\subset U$, i.e.
$$U^0:=\bigcap_{u\in U} (U^1)^u\ .$$
\end{ddd}
$U^0\subset U$ has finite index, too.
The torus $M_{U^0}$ coincides with $\overline{m(N_V)}$.
\begin{ddd}\label{weih1422}
We set $P_{U^0}:=N_VM_{U^0}$. Furthermore we define
$M_U:=\overline{l_M(U)}$ and $P_U:=N_VM_U$.
\end{ddd}
Then $U\subset P_U$, $M_{U^0}\subset M_U$ has finite index, and $M_U$
normalizes $N_V$.
\subsubsection{}
We have an exact sequence
$$0\rightarrow E\rightarrow P\stackrel{l_A}{\rightarrow} A\rightarrow 0 \ ,$$
where the decomposition $E=NM$ is already fixed by the split $s$.
The split extends uniquely to a split
$s:MA\rightarrow P$ of (\ref{yy77}) and induces a split
of the sequence above.
In particular we
obtain a Langlands decomposition $P=NAM$, where
we identify $A$ with its image by $s$. The split
furthermore defines an action of $A$ on $N$ by
automorphisms commuting with the action of $M$.
\subsubsection{}
We call the cusp associated to $U\subset P$ regular
if we can choose $s$ such that $N_V$ is $A$-invariant (compare Definition \ref{t799}).
\begin{lem}\label{martin}
If $X$ is a real or complex hyperbolic space,
then every cusp is regular.
\end{lem}
{\it Proof.$\:\:\:\:$}
If $X$ is real-hyperbolic, then for any split $s$ the group $A$
acts on ${\mathfrak{n}}$ as multiplication by scalars. Any subspace,
in particular ${\mathfrak{n}}_V$, is invariant with respect to $A$.
We now consider the case that $X={H\mathbb{C}}^n$.
Assume that we have chosen a split $s$. Then we can decompose
${\mathfrak{n}}={\mathfrak{n}}_{\alpha}\oplus{\mathfrak{n}}_{2\alpha}$ with respect to the action
of $A$ such that $a\in A$ acts on ${\mathfrak{n}}_{i\alpha}$ as multiplication
by $a^{i\alpha}$. Here ${\mathfrak{n}}_{\alpha}$ is a symplectic
vector space of dimension $2(n-2)$,
where the symplectic form
with values in the one-dimensional space ${\mathfrak{n}}_{2\alpha}$
is given by the commutator.
Taking $M_{U^0}$-invariants we obtain a decomposition
${\mathfrak{n}}^{M_{U^0}}={\mathfrak{n}}_1\oplus{\mathfrak{n}}_2$,
where ${\mathfrak{n}}_1\subset {\mathfrak{n}}_{\alpha}$ is a
symplectic subspace (\cite{guilleminsternberg77}, Prop. 4.2.1). If ${\mathfrak{n}}_1=\{0\}$,
then ${\mathfrak{n}}_V={\mathfrak{n}}_2$ is invariant with respect to $A$,
and we are done. We now assume that
${\mathfrak{n}}_1\not=\{0\}$.
We have two cases. If ${\mathfrak{n}}_2\subset {\mathfrak{n}}_V$,
then ${\mathfrak{n}}_V={\mathfrak{n}}_V\cap{\mathfrak{n}}_1\oplus {\mathfrak{n}}_2$,
and this space is invariant with respect to $A$.
Thus we assume that ${\mathfrak{n}}_2\not\subset{\mathfrak{n}}_V$.
Since $\dim({\mathfrak{n}}_2)=1$ we then have ${\mathfrak{n}}_2\cap {\mathfrak{n}}_V=\{0\}$.
Let $p_i:{\mathfrak{n}}_V\rightarrow {\mathfrak{n}}_i$ denote the projections.
Then $p_1:{\mathfrak{n}}_V\rightarrow {\mathfrak{n}}_1$ is injective.
Let $\lambda:{\mathfrak{n}}_1\rightarrow {\mathfrak{n}}_2$ be a linear extension
of $p_2\circ p_1^{-1}:p_1({\mathfrak{n}}_V)\rightarrow {\mathfrak{n}}_2$.
Since the symplectic form is non-degenerate
there exists a unique $Y\in {\mathfrak{n}}_1$ such that
$\lambda(X)=[Y,X]$ for all $X\in{\mathfrak{n}}_1$.
Let $h:=\exp(-Y)\in N^{M_{U^0}}$.
We claim that ${\mathfrak{n}}_V^h\subset {\mathfrak{n}}_1$.
Indeed, if $X\in {\mathfrak{n}}_V$,
then $p_2(X^h)=p_2(X-[Y,X])=\lambda(p_1(X))-[Y,X]=[Y,X]-[Y,X]=0$.
If we replace the split $s:M\times A\rightarrow P$
by $hsh^{-1}$, then ${\mathfrak{n}}_V$ gets replaced by ${\mathfrak{n}}_V^h$.
Thus by an appropriate choice of the split
we can assume that ${\mathfrak{n}}_V\subset{\mathfrak{n}}_1$, and we are done.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
The next lemma shows that regularity of cusps is a proper restriction in the case of ${H\mathbb{H}}^n$.
\begin{lem}\label{contrmartin}
If $X={H\mathbb{H}}^n$, $n\ge 2$, then there exist non-regular
cusps.
\end{lem}
{\it Proof.$\:\:\:\:$}
It suffices to provide an example for $n=2$.
We can identify
${\mathfrak{n}}={\mathbb{H}}\oplus {\rm Im}_{\mathbb{H}}({\mathbb{H}})$ (${\rm Im}_{\mathbb{H}}({\mathbb{H}})$ denoting the imaginary quaternions),
and the commutator
of $X,Y\in{\mathbb{H}}$ is given by
$[X,Y]= \bar X Y-\bar Y X \in {\rm Im}_{\mathbb{H}}({\mathbb{H}})$.
Implicitly we have fixed some split $s\times t$
such that $a\in A$ acts on ${\mathbb{H}}$ as multiplication by
$a^\alpha$ and on ${\rm Im}_{\mathbb{H}}({\mathbb{H}})$ by $a^{2\alpha}$.
Let $1,I,J,K$ be a base of the copy of ${\mathbb{H}}$ and $i,j,k$
be the base of ${\rm Im}_{\mathbb{H}}({\mathbb{H}})$.
Then we consider ${\mathfrak{n}}_V={\mathrm{span}}_{\mathbb{R}}\{1,I+j,i\}$
and let $U\subset N_V$ be any lattice.
It is not possible to conjugate this subspace
into an $A$-invariant one.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Assume that $U\subset P$ defines a regular cusp.
\begin{ddd}\label{weih215}
We define $\rho_U\in{\mathfrak{a}}^*$ by $\rho_U(H):=\frac12 {\mbox{\rm tr}}({\mbox{\rm ad}}(H)_{|{\mathfrak{n}}_V})$,
where ${\mathfrak{n}}_V$ is the Lie algebra of $N_V$. Furthermore, we set
$\rho^U:=\rho-\rho_U$.
\end{ddd}
\subsection{Schwartz spaces}\label{schw}
\subsubsection{}
In this subsection we start the description of our function spaces. Here we introduce the Schwartz spaces ${S}_{U}(\sigma_\lambda,{\varphi})$. We will show that these spaces form trivial holomophic bundles of Fr\'echet spaces for $\lambda\in {\aaaa_\C^\ast}$, and that they are compatible with twisting. The Schwartz spaces are basic building blocks of the function spaces $B_U(\sigma_\lambda,{\varphi})$ which we will define later
(see Subsection \ref{asz}) .
\subsubsection{}
We fix a representative $w\in K$ of the non-trivial element of the Weyl group $W({\mathfrak{g}},{\mathfrak{a}})$ such that $w^2=1$. Let $\infty_P\in\partial X$
be the fixed point of $P$ and $0_P:=w\infty_P$.
The subset
$\Omega_U=\Omega_P=\partial X\setminus\infty_P$
is the $N$-orbit of $0_P$ (see Subsection \ref{geomf} for notation). Indeed, the map
$N\ni x\mapsto xw\infty_P\in \Omega_U$ is a diffeomorphism from $N$ to $\Omega_U$.
\subsubsection{}
Let $(\sigma,V_\sigma)$ be a finite-dimensional unitary representation
of $M$. Given $\lambda\in {\aaaa_\C^\ast}$ we form the representation $(\sigma_\lambda,V_{\sigma_\lambda})$ of $P$ (see \ref{weih135}). Furthermore,
let $({\varphi},V_{\varphi})$ be an admissible twist (see Definition \ref{weih134}).
\subsubsection{}\label{weih243}
The action of $A$ on ${\mathcal{U}}({\mathfrak{n}})$ induces a grading.
We choose a basis $\{A_i\}$ of ${\mathcal{U}}({\mathfrak{n}})$ such that the subset
$\{A_i\:|\: \deg(A_i)\le d\alpha\}$ spans ${\mathcal{U}}({\mathfrak{n}})^{\le d\alpha}$.
Furthermore, we choose a norm $|.|$ on $V_{\varphi}$.
For any $d\in{\mathbb {N}}_0$, $k\in {\mathbb{R}}$, and compact subset $W\subset N\setminus N_V$ (see Theorem \ref{ausl} for the definition of $N_V$) we define
the seminorm $q_{W,d,k}$ on $C_c^\infty(B_{U},V_{B_{U}}(\sigma_\lambda,{\varphi}))$
by
$$q_{W,d,k}(f):=\sup_{\{i|\deg(A_i)\le d\alpha\}} \sup_{x\in W}\sup_{a\in A_+}
a^{k\alpha-2(\lambda-\rho^U)} |{\varphi}(a)^{-1} f(x^aA_iw)|\ .$$
Here we consider a section of $V_{B_{U}}(\sigma_\lambda,{\varphi})$ as a
$U$-invariant function on $G\setminus P$ with values in $V_{\sigma_\lambda}\otimes V_{\varphi}$
satisfying the corresponding invariance conditions with respect to the
right $P$-action, and where $u\in U$ acts by
$(\pi^{\sigma_\lambda,{\varphi}}(u)f)(xw)={\varphi}(u) f(u^{-1}xw)$.
\subsubsection{}\label{weih259}
We fix $W$ and another compact subset $W^\prime\subset N$ such that
$W^\prime\cup W^{A_+}$ projects surjectively onto $B_{U}$.
We add an arbitrary $C^d$-norm over $W^\prime$ to $q_{W,d,k}$
in order to obtain a norm $\|.\|_{k,d}$.
We first define $S_{U,k,d}(\sigma_\lambda,{\varphi})$ to be
the Banach space closure of $C_c^\infty(B_{U},V_{B_{U}}(\sigma_\lambda,{\varphi}))$
with respect to $\|.\|_{k,d}$. This Banach space is independent
(up to equivalent norms)
of the choices of $W$, $W^\prime$, $|.|$, and the base of ${\mathcal{U}}({\mathfrak{n}})$.
\begin{ddd}\label{weih136}
We define ${S}_{U,k}(\sigma_\lambda,{\varphi})$ to be the intersection of
the spaces ${S}_{U,k,d}(\sigma_\lambda,{\varphi})$ for all $d\in{\mathbb {N}}_0$.
The space ${S}_{U,k}(\sigma_\lambda,{\varphi})$ is a Fr\'echet space which
is topologized by the countable set of norms $\|.\|_{k,d}$, $d\in{\mathbb {N}}_0$.
\end{ddd}
The space ${S}_{U,k}(\sigma_\lambda,{\varphi})$ is a space of smooth sections with a fixed growth rate at infinity
of $B_U$ measured by $k\in {\mathbb {N}}_0$.
\subsubsection{}
We now show that the family of spaces $\{{S}_{U,k}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$
forms a trivial holomorphic bundle of Fr\'echet spaces. To this end we construct a certain holomorphic family of functions $s^z$, $z\in {\mathbb {C}}$. In the proof of Lemma \ref{ytra} we use multiplication by these functions in order to identify
the Schwartz spaces for different $\lambda\in {\aaaa_\C^\ast}$.
\subsubsection{}
\begin{lem}\label{ingf}
There exists a positive $P_{U}$-invariant function
$s\in C^\infty(N)$ such
that for any compact subset $W\subset N\setminus N_V$ there is $a_0\in A$ with
$s(x^a)=a^{2\alpha}s(x)$ for all $a\ge a_0$ and $x\in W$.
\end{lem}
{\it Proof.$\:\:\:\:$}
Let $\bar{N}:=N^w$ and consider the decomposition
$$G\setminus wP=\bar NMAN\ , \quad g=\bar n(g) m(x)a(x)n(x)\ .$$
Note that $a(xw)$ satisfies $a(x^aw)=a^2a(xw)$ for all $a\in A$ and $x\in N\setminus\{1\}$ (see (\ref{weih141}) below).
We consider the function $N\setminus\{1\}\ni x\mapsto a(xw)^{\alpha}$ as a positive
function $s_1\in C^\infty(N\setminus \{1\})$.
Let $s_2\in C^\infty(N)$ be any positive function.
Using a partition of unity we glue $s_1$ and $s_2$
to obtain a positive function $s_3\in C^\infty(N)$
which coincides with $s_1$ outside of a compact subset $W_1$ of
$N$. We now choose $z\in{\mathbb{R}}$ such $z<-\rho_U/\alpha$
and define $s_4$ as the average
$$s_4(x):=\int_{P_{U}} s_3^z(y.x) dy$$
(see Definition \ref{weih1422} for $P_U$ and Theorem \ref{ausl}, (2) for the action $(y,x)\mapsto y.x$).
The integral converges (compare Lemma \ref{210}) and defines a positive $P_{U}$-invariant smooth function on $N$.
If $W\subset N\setminus N_V$ is compact, then there is $a^0\in A$
such that $(N_VW)^{aM_U}\cap W_1=\emptyset$ for all $a\ge a_0$.
If $a\ge a_0$ and $x\in W$, then we have
\begin{eqnarray*}
s_4(x^a)&=&\int_{P_{U}} s_3^z(y.x^a) dy\\
&=&\int_{M_U}\int_{N_V} s_3^z(vx^{au}) dv du\\
&=&a^{2\rho_U}\int_{M_U}\int_{N_V} s^z_3((vx)^{au}) dvdu\\
&=&a^{2(z\alpha+\rho_U)} s_4(x)\ .
\end{eqnarray*}
We define $s$ as the $\alpha/(z\alpha+\rho_U)$'th power of $s_4$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}\label{neuj911}
We now consider the function $s$ constructed in Lemma \ref{ingf}
as a section $s\in C^\infty(\Omega_P,V(1_{\rho+\alpha}))$
by defining $s(xw):=s(x)$.
If $\lambda\in{\mathfrak{a}}^*$, then the bundle $V(1_\lambda)$ is a
complex line bundle which can be written as the complexification
of a trivial $G$-equivariant real line bundle $V(1_\lambda)^{\mathbb{R}}$.
It makes sense to speak of a positive
section of $V(1_\lambda)^{\mathbb{R}}$.
A section of $V(1_\lambda)$ is called positive if it is
a positive section of the real subbundle $V(1_\lambda)^{\mathbb{R}}$.
In particular, the section $s$ of $V(1_{\rho+\alpha})$
constructed above is positive.
\subsubsection{}\label{neuj910}
For each $z\in {\mathbb {C}}$ we can form $s^z$, which is a
section of $V(1_{\rho+z\alpha})$. There is a natural
identification $V(\sigma_\lambda)\otimes V(1_{\rho+z\alpha})\cong V(\sigma_{\lambda+z\alpha})$.
Multiplication by $s^z$
identifies $C^\infty(\Omega_P,V(\sigma_\lambda,{\varphi}))$ with $C^\infty(\Omega_P,V(\sigma_{\lambda+z\alpha},{\varphi}))$.
\begin{lem}\label{ytra}
For each $d\in{\mathbb {N}}_0$ and $z\in{\mathbb {C}}$
multiplication by $s^z$ defines a continuous map from
${S}_{U,k,d}(\sigma_\lambda,{\varphi})$ to ${S}_{U,k,d}(\sigma_{\lambda+z\alpha},{\varphi})$.
The family
$\{{S}_{U,k}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$ is a
trivial holomorphic bundle of Fr\'echet spaces.
\end{lem}
{\it Proof.$\:\:\:\:$}
Since $s$ is $U$-invariant and non-vanishing multiplication by $s^z$
is an isomorphism of $C^\infty(B_{U},V_{B_{U}}(\sigma_\lambda,{\varphi}))$
with $C^\infty(B_{U},V_{B_{U}}(\sigma_{\lambda+z\alpha},{\varphi}))$
and of the subspaces of sections with compact support.
It suffices to show that there
is a constant $C\in {\mathbb{R}}$ such that
for all $f\in C_c^\infty(B_{U},V_{B_{U}}(\sigma_\lambda,{\varphi}))$
we have $\|s^z f\|_{k,d}\le C \|f\|_{k,d}$.
This follows from the Leibniz rule and the following estimate.
Let $A\in{\mathcal{U}}({\mathfrak{n}})$ be homogeneous of degree $d\alpha$. Then for any compact subset
$W\subset N\setminus N_V$ there is a constant $C\in{\mathbb{R}}$ and $a_0\in A$
such that for $x\in W$ and $a\ge a_0$
\begin{eqnarray*}
|s^z(x^aAw)|&=&|s^z((xA^{a^{-1}})^aw)|\\
&=&a^{(2z-d)\alpha} |s^z(xAw)|\\
&\le& C a^{(2z-d)\alpha} \ .
\end{eqnarray*}
For any $\lambda_0$ we now define the trivialization
$\Phi_{\lambda_0}:\bigcup_{\lambda\in{\aaaa_\C^\ast}} {S}_{U,k}(\sigma_{\lambda},{\varphi})\rightarrow {S}_{U,k}(\sigma_{\lambda_0},{\varphi}) \times {\aaaa_\C^\ast}$
such that the restriction of $\Phi_{\lambda_0}$ to the fibre ${S}_{U,k}(\sigma_{\lambda},{\varphi})$
is multiplication by $s^{(\lambda_0-\lambda)/\alpha}$.
The transition map $\Phi_{\lambda_0}\circ \Phi_{\lambda_1}^{-1}$
is multiplication by $s^{(\lambda_0-\lambda_1)/\alpha}$
and thus independent of $\lambda$. In particular, it
is a holomorphic family of continuous maps.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
The Schwartz space is the space of smooth rapidly decaying sections on $B_U$.
\begin{ddd}\label{weih245}
We define the Schwartz space
$S_{U}(\sigma_\lambda,{\varphi})$ as the intersection of the spaces
$S_{U,k}(\sigma_\lambda,{\varphi})$ over all $k\in{\mathbb {N}}_0$.
\end{ddd}
\subsubsection{}
Since the trivializations $\Phi_\lambda$ are compatible with the
inclusions $S_{U,k^\prime}(\sigma_\lambda,{\varphi})\hookrightarrow
S_{U,k}(\sigma_\lambda,{\varphi})$, $k^\prime\ge k$, we obtain
the following corollary.
\begin{kor}
The family $\{S_{U}(\sigma_\lambda,{\varphi})\}_{\lambda\in {\aaaa_\C^\ast}}$
is a trivial holomorphic vector bundle of Fr\'echet spaces.
\end{kor}
\subsubsection{}
Note that the inclusions $S_{U,k+1}(\sigma_\lambda,{\varphi})\hookrightarrow S_{U,k}(\sigma_\lambda,{\varphi})$ are compact. The proof of this fact is a simple application of the Lemma of Arzela-Ascoli.
The compactness of these inclusions imply the following fact.
\begin{kor}
The Schwartz space
$S_{U}(\sigma_\lambda,{\varphi})$ is a Montel space. In particular, it is reflexive.
\end{kor}
\subsubsection{}
If the cusp associated to $U\subset P$
has full rank, then we have for all $k\in{\mathbb {N}}_0$
$$S_{U,k}(\sigma_\lambda,{\varphi})={S}_{U}(\sigma_\lambda,{\varphi})=C^\infty(B_{U},V_{B_{U}}(\sigma_\lambda,{\varphi}))\ .$$
\subsubsection{}
Next we show that the Schwartz spaces are compatible
with twisting.
Let $\sigma$ be a Weyl invariant representation of $M$ and
$(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$ be a finite-dimensional
representation of $G$ as in \ref{weih144}. See \ref{weih146} for the definition of $Z(\lambda)$ and $\Pi(\lambda)$.
\begin{lem}\label{compat1}
\mbox{}\\
\begin{enumerate}
\item The Schwartz space
${S}_{U}(\sigma_\lambda,{\varphi})$ coincides with the
closed subspace of ${}^U C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$
of sections which vanish at $\infty_P$ of infinite order.
\item
We have holomorphic families of continuous maps
\begin{eqnarray*}
\{i_{\sigma,\mu}^U\}&:&S_{U}(\sigma_\lambda,{\varphi})\rightarrow
S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})\\
\{p_{\sigma,\mu}^U\}&:&S_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow
{S}_{U}(\sigma_\lambda,{\varphi})\ .
\end{eqnarray*}
\item
$S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
is a $Z({\mathfrak{g}})$-module.
\item
If $\lambda\not\in I_{\mathfrak{a}}$, then
$\{i_{\sigma,\mu}^U\}$ maps ${S}_{U}(\sigma_\lambda,{\varphi})$
isomorphically onto $$\ker\{Z(\lambda):S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})\}\ ,$$
and the restriction of
$\{p_{\sigma,\mu}^U\}$ to $$\ker\{\tilde Z(\lambda):S_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow S_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\}$$
is an isomorphism onto ${S}_{U}(\sigma_\lambda,{\varphi})$
(here $\tilde Z(\lambda)$ is the adjoint of
$\pi^{1_{-\lambda+\mu}\otimes \tilde\pi_{\sigma,\mu},{\mathrm {id}}_{{\varphi}}}(\Pi(\lambda))$).
\item
The composition
$$\{j_{\sigma,\mu}^U\}:S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})
\stackrel{1-Z(\lambda)}{\rightarrow} \ker(Z(\lambda))\stackrel{(\{i_{\sigma,\mu}^U\})^{-1}}{\rightarrow}
{S}_{U}(\sigma_\lambda,{\varphi})$$
which is intitially defined for $\lambda\not\in I_{\mathfrak{a}}$
extends to a meromorphic family of continuous maps.
Similarly, the composition
$$\{q_{\sigma,\mu}^U\}:{S}_{U}(\sigma_\lambda,{\varphi})\stackrel{(\{p_{\sigma,\mu}^U\})^{-1}}{\rightarrow} \ker(\tilde Z(\lambda)) \rightarrow S_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})$$
extends to a meromorphic family of continuous maps.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
We leave the proof of 1. at this place to the interested reader.
Alternatively the assertion can be considered as an immediate consequence
of the material of Subsection \ref{trivas}.
\subsubsection{}
2. follows from 1. and the fact that $\{i_{\sigma,\mu}^U\}$
and $\{p_{\sigma,\mu}^U\}$ are induced by an inclusion
$$i_{\sigma,\mu}:V(\sigma_\lambda,{\varphi})\rightarrow V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$$
and a projection $$p_{\sigma,\mu}:V(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow V(\sigma_\lambda,{\varphi})$$
of bundles. Another argument for the assertion about
$\{i_{\sigma,\mu}^U\}$ would be to check that
$\{i_{\sigma,\mu}^U\}$ maps $S_{U,k,d}(\sigma_\lambda,{\varphi})$ continuously to
$S_{U,k^\prime,d}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$,
$k^\prime=k+\mu/\alpha$,
for all $k,d\in{\mathbb {N}}_0$ by comparing the norms explicitly.
A similar argument works for $\{p_{\sigma,\mu}^U\}$ as well.
\subsubsection{}
In order to see 3. note that the action of $cZ({\mathfrak{g}})$ is implemented
by differential operators. Using the fact that the action
$\pi^{1_{\lambda+\mu}\otimes \pi_{\sigma,\mu},{\mathrm {id}}_{\varphi}}$
of ${\mathcal{Z}}({\mathfrak{g}})$ commutes with the action
$\pi^{1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi}}$
of $P_UA$ which is used in order to characterize
the growth of elements of $S_{U,k}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
one could alternatively check that this space is a ${\mathcal{Z}}({\mathfrak{g}})$-module.
\subsubsection{}
We now prove the first
assertion of 4. and leave the second to the reader since the argument
is similar. The representation $V_{1_{\lambda+\mu}}\otimes V_{\pi_{\sigma,\mu}}$ of $P$ fits
into an exact sequence
$$0\rightarrow V_{\sigma_\lambda} \stackrel{i_{\sigma,\mu}}{\rightarrow} V_{1_{\lambda+\mu}}\otimes V_{\pi_{\sigma,\mu}}
\rightarrow W\rightarrow 0\ ,$$
which induces a corresponding exact sequence of bundles
\begin{equation}
\label{hj112}0\rightarrow V(\sigma_\lambda,{\varphi})\stackrel{i_{\sigma,\mu}}{\rightarrow} V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})
\rightarrow V(w ,{\varphi}) \rightarrow 0 \ ,\end{equation}
where $w $ denotes the representation of $P$ on $W $. Using 1. we obtain an exact
sequence of sections which vanish of infinite order at $\infty_P$.
Going over to $U$-invariant sections we obtain
the exact sequence
$$0\rightarrow S_U(\sigma_\lambda,{\varphi})\stackrel{\{i_{\sigma,\mu}^U\}}{\rightarrow} S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})
\stackrel{p}{\rightarrow} S_U(w ,{\varphi}) \ .$$
The action of ${\mathcal{Z}}({\mathfrak{g}})$ commutes with $p$.
The operator $\Pi(\lambda)$ is designed such that
$\pi^{w ,{\varphi}}(\Pi(\lambda))=1$
and $\pi^{\sigma_\lambda,{\varphi}}(\Pi(\lambda))=0$ (note that $\lambda\not\in I_{\mathfrak{a}}$).
Since $\{i_{\sigma,\mu}^U\}$ is a morphism of $Z({\mathfrak{g}})$-modules
we see that $\{i_{\sigma,\mu}^U\}$ maps to $\ker(Z(\lambda))$.
In order to show that it is onto take $f\in \ker(Z(\lambda))$.
Then $$p(f)=p((1-Z(\lambda))f)=(1-\pi^{w,{\varphi}}(\Pi(\lambda)))p(f)=0\ ,$$
and $f$ is in the range of $\{i_{\sigma,\mu}^U\}$.
\subsubsection{}
Finally we show the first assertion of 5. and leave the second to the reader
since the argument is similar, again.
We choose a holomorphic family of splits $j :V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}) \rightarrow
V(\sigma_\lambda,{\varphi})$ of the sequence of bundles (\ref{hj112}) (not necessarily $U$-invariant). The split
$j $ induces a holomorphic family of maps $J :C^\infty(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\rightarrow
C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$.
The map $J$ is compatible with the subspaces of sections
vanishing of infinite order at $\infty_P$. We define
the meromorphic family of continuous maps
$\tilde j_{\sigma,\mu}^U$ as the restricition of $J \circ (1-Z(\lambda))$
to $S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$.
Since $$\tilde j_{\sigma,\mu}^U\circ i_{\sigma,\mu}^U = J \circ (1-Z(\lambda))\circ i_{\sigma,\mu}^U={\mathrm {id}}$$
and the restriction of $\tilde j_{\sigma,\mu}^U$
to $\ker(1-Z(\lambda))$ vanishes we see by 4. that
$\tilde j_{\sigma,\mu}^U$ maps to $S_U(\sigma_\lambda,{\varphi})$,
and that it coincides with $\{j_{\sigma,\mu}^U\}$ for $\lambda\not\in I_{\mathfrak{a}}$.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsection{Asymptotics for the trivial group}\label{trivas}
\subsubsection{}\label{weih600}
We consider the decomposition $G\setminus wP=\bar{N}MAN$,
$g=\bar{n}(g)m(g)a(g)n(g)$. There is a unique diffeomorphism
$F:N\setminus \{1\} \rightarrow \bar{N}\setminus \{1\}$
such that $nw\in F(n)MAN$. Indeed, $F(n)=\bar{n}(nw)$. One can check that
\begin{eqnarray}
a(n^{ma}w)&=&a^2 a(nw)\label{weih141}\\
m(n^aw)&=&m(nw)\nonumber\\
F(n^{ma})&=& F(n)^{ma}\nonumber\ .
\end{eqnarray}
\subsubsection{}\label{weih213}
Let $({\varphi},V_{\varphi})$ be an admissible twist for $U$.
For simplicity we normalize the restriction of ${\varphi}$
to $A$ such that
the lowest $A$-weights of all its irreducible coponents are zero.
Let $l_{\varphi}\in{\mathfrak{a}}^*$ be the highest weight
of ${\varphi}$.
\subsubsection{}
In the present subsection we describe the space
$C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))=:B_{\{1\}}(\sigma_\lambda,{\varphi})$
as an extension
$$0\rightarrow {S}_{\{1\}}(\sigma_\lambda,{\varphi})\rightarrow B_{\{1\}}(\sigma_\lambda,{\varphi})
\rightarrow {R}_{\{1\}}(\sigma_\lambda,{\varphi})\rightarrow 0\ .$$
The Schwartz space ${S}_{\{1\}}(\sigma_\lambda,{\varphi})$
coincides with the subspace of $C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$ of all
section that vanish at $\infty_P$ of infinite order.
Note that $\bar N$ is diffeomorphic to $\partial X\setminus 0_P$
by $\bar n\mapsto \bar n \infty_P$
(recall that $0_P=w\infty_P$).
Using a theorem
of E. Borel to the effect that each formal power series
can be realized as a Taylor series of a smooth function we obtain
an exact sequence
\begin{equation}\label{borell}0\rightarrow S_{\{1\}}(\sigma_\lambda,{\varphi})\rightarrow C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))
\stackrel{TS}{\rightarrow} {\mbox{\rm Hom}}({\mathcal{U}}(\bar{\mathfrak{n}}),V_{\sigma_\lambda}\otimes V_{\varphi}) \rightarrow 0\ ,\end{equation}
where $TS$ is given by $TS(f)(A):=f(Ae)$.
\subsubsection{}\label{neuj100}
The space ${\mbox{\rm Hom}}({\mathcal{U}}(\bar{\mathfrak{n}}),V_{\sigma_\lambda}\otimes V_{\varphi})$ admits
an action of $AM_U$ by $(u.f)(A):=(\sigma_\lambda(u)\otimes {\varphi}(u))f(A^{u^{-1}})$.
With respect to the action of $A$ it
can be decomposed as
$${\mbox{\rm Hom}}({\mathcal{U}}(\bar{\mathfrak{n}}),V_{\sigma_\lambda}\otimes V_{\varphi})=\prod_{n\in{\mathbb {N}}_0} {\mbox{\rm Hom}}({\mathcal{U}}(\bar{\mathfrak{n}}),V_{\sigma_\lambda}\otimes V_{\varphi})^n\ ,$$
where $A$ acts on the $M_U$-module ${\mbox{\rm Hom}}({\mathcal{U}}(\bar{\mathfrak{n}}),V_{\sigma_\lambda}\otimes V_{\varphi})^n$
with weight $\rho-\lambda+\alpha n$.
\subsubsection{}
By definition a function $p:\bar N\rightarrow V_{\sigma_\lambda}\otimes V_{\varphi}$ is a polynomial
iff $p\circ \exp:\bar{\mathfrak{n}}\rightarrow V_{\sigma_\lambda}\otimes V_{\varphi}$ is a polynomial.
Such a function
$p$ is called homogeneous of degree $n$, iff
$$(\sigma_{\lambda}(a)\otimes {\varphi}(a))p(\bar n^{a^{-1}})= a^{\rho-\lambda+n\alpha} p(\bar n)\ .$$
\subsubsection{}\label{neuj101}
The space ${\mbox{\rm Hom}}({\mathcal{U}}(\bar{\mathfrak{n}}),V_{\sigma_\lambda}\otimes V_{\varphi})^n$ can be identified
$AM_U$-equivariantly with the space of polynomials ${\rm Pol}(\bar N,V_{\sigma_\lambda}\otimes V_{\varphi})^n$
on $\bar N$ with values in $V_{\sigma_\lambda}\otimes V_{\varphi}$ being homogeneous of degree $n$.
Here a polynomial $p\in {\rm Pol}(\bar N,V_{\sigma_\lambda}\otimes V_{\varphi})^n$
corresponds to the map ${\mathcal{U}}(\bar{\mathfrak{n}})\ni A\mapsto
p(Ae)\in V_{\sigma_\lambda}\otimes V_{\varphi}$.
\subsubsection{}\label{weih503}
If a homogeneous polynomial $p\in {\rm Pol}(\bar N,V_{\sigma_\lambda}\otimes V_{\varphi})^n$
is considered as a section $f_p$ of $V(\sigma_\lambda,{\varphi})$ defined over
$\partial X\setminus \{0_P\}$, then we have for $x\in N$
$$f_p(xw)=f_p(F(x)m(xw)a(xw)n(xw))=\sigma(m(xw))^{-1} a(xw)^{\lambda-\rho} f_p(F(x))\ .$$
In particular, for $x\in N$, $x\not=1$, and $a\in A$ we have
\begin{eqnarray*}
{\varphi}(a)^{-1} f_p(x^aw)&=&{\varphi}(a)^{-1}a(x^aw)^{\lambda-\rho} \sigma(m(x^aw))^{-1}f_p(F(x^a))\\
&=&
a^{2(\lambda-\rho)-n\alpha} a(xw)^{\lambda-\rho}\sigma(m(xw))^{-1} f_p(F(x))\ .
\end{eqnarray*}
\begin{ddd}\label{weih231}
We define $A_{\{1\}}(\sigma_\lambda,{\varphi})^n\subset C^\infty(\partial X\setminus
\{\infty_P,0_P\},V(\sigma_\lambda,{\varphi}))$ as the subspace
of sections $f$ satisfying
\begin{equation}\label{weih217}
{\varphi}(a)^{-1}f(x^aw)=a^{2(\lambda-\rho)-n\alpha} f(xw)
\end{equation}
for all $x\in N\setminus \{1\}$, $a\in A$.
We define the subspace ${R}_{\{1\}}(\sigma_\lambda,{\varphi})^n\subset A_{\{1\}}(\sigma_\lambda,{\varphi})^n$
as the subspace spanned by the sections $f_p$ corresponding
to $p\in{\rm Pol}(\bar N,V_{\varphi})^n$.
We further define ${R}_{\{1\},k}(\sigma_\lambda,{\varphi}):=\bigoplus_{n=0}^k {R}_{\{1\}}(\sigma_\lambda,{\varphi})^n$ and ${R}_{\{1\}}(\sigma_\lambda,{\varphi}):=\prod_{n=0}^\infty{R}_{\{1\}}(\sigma_\lambda,{\varphi})^n$.
\end{ddd}
\subsubsection{}
The spaces $A_{\{1\}}(\sigma_\lambda,{\varphi})^n$ for the various $\lambda$
can be identified using multiplication by suitable powers of
the function $xw\mapsto a(xw)^\alpha$ and therefore form a trivial
vector bundle of Fr\'echet spaces over ${\aaaa_\C^\ast}$. The spaces
${R}_{\{1\}}(\sigma_\lambda,{\varphi})^n$ form trivial finite-dimensional
subbundles trivialized by the identification with the trivial bundles
${\aaaa_\C^\ast}\times {\rm Pol}(\bar N,V_{\sigma} \otimes V_{\varphi})^n\rightarrow {\aaaa_\C^\ast}$.
Thus ${R}_{\{1\},k}(\sigma_\lambda,{\varphi})$ has the structure of a trivial
holomorphic vector bundle.
\subsubsection{}\label{neuj202}
We fix a smooth cut-off function $\chi\in C^\infty(A)$ such that
$\chi(a)=0$ in a neighbourhood of $A_-$ and $\chi(a)=1$ if $a^\alpha>2$.
Multiplication by the function
$$\Omega_P\ni x\,0_P\mapsto \chi(a(xw))$$
defines inclusions
$L: {R}_{\{1\}}(\sigma_\lambda,{\varphi})^n \rightarrow C^\infty(B_{\{1\}},V_{\{1\}}(\sigma_\lambda,{\varphi}))$
for each $n$, and summing these maps up we obtain inclusions
$$L: {R}_{\{1\},k}(\sigma_\lambda,{\varphi}) \rightarrow C^\infty(B_{\{1\}},V_{B_{\{1\}}}(\sigma_\lambda,{\varphi}))\ .$$
In the present paper the symbol $L$ is used for various maps of this kind. It will always be clear from the context which version of $L$ is meant.
\subsubsection{}
If $f\in {S}_{\{1\},k}(\sigma_\lambda,{\varphi})$,
then
$$\lim_{a\to\infty} a^{k\alpha-2(\lambda-\rho)}{\varphi}(a)^{-1} f(x^aw) =0$$
uniformly for $x$ in compact subsets of $N\setminus \{1\}$.
Thus $L({R}_{\{1\},k}(\sigma_\lambda,{\varphi}))$ intersects ${S}_{\{1\},k}(\sigma_\lambda,{\varphi})$ trivially. We define
$$B_{\{1\},k}(\sigma_\lambda,{\varphi}):={S}_{\{1\},k}(\sigma_\lambda,{\varphi})\oplus L({R}_{\{1\},k}(\sigma_\lambda,{\varphi}))\ .$$
This space fits into the exact sequence
\begin{equation}\label{weih210}
0\rightarrow {S}_{\{1\},k}(\sigma_\lambda,{\varphi}) \rightarrow B_{\{1\},k}(\sigma_\lambda,{\varphi})
\stackrel{AS}{\rightarrow}{R}_{\{1\},k}(\sigma_\lambda,{\varphi})\rightarrow 0\ ,
\end{equation}
where $AS$ takes the finite asymptotic expansion. To $AS$ applies the same remark as as above for $L$. This symbol appears in various versions which will be denoted by the same symbol, and it will be clear from the context which version is meant.
\subsubsection{}
The sequence (\ref{weih210}) is split by $L$
and defines $B_{\{1\},k}(\sigma_\lambda,{\varphi})$ as a topological vector space
and the family of spaces $\{B_{\{1\},k}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$
as a trivial bundle of Fr\'echet spaces.
Moreover, we have continuous and compact inclusions
$B_{\{1\},k+1}(\sigma_\lambda,{\varphi})\hookrightarrow B_{\{1\},k}(\sigma_\lambda,{\varphi})$.
The intersection of these spaces over all $k\in{\mathbb {N}}_0$ is the Fr\'echet
and Montel space $B_{\{1\}}(\sigma_\lambda,{\varphi})$, which concides
with $C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$ because of
exactness of (\ref{borell}).
Note that the
sequence
$$0\rightarrow {S}_{\{1\}}(\sigma_\lambda,{\varphi})\rightarrow B_{\{1\}}(\sigma_\lambda,{\varphi})
\rightarrow {R}_{\{1\}}(\sigma_\lambda,{\varphi})\rightarrow 0$$
does not admit any continuous split.
\subsubsection{}
Though this is clearly possible we will not attempt to
trivialize the family of spaces $B_{\{1\}}(\sigma_\lambda,{\varphi})$.
But we keep in mind that $B_{\{1\}}(\sigma_\lambda,{\varphi})$ is an intersection
of trivial holomorphic bundles (see \ref{weih211}).
Below we will obtain a similar description of the spaces
$B_{U,k}(\sigma_\lambda,{\varphi})$.
\subsection{The spaces $B_{U,k}(\sigma_\lambda,{\varphi})$}\label{asz}
\subsubsection{}
In the present subsection $\sigma$ is any finite-dimensional representation of $M$, and ${\varphi}$ denotes a twist. Below we define the spaces $B_{U}(\sigma_\lambda,{\varphi})$.
In particular, we construct the spaces
${R}_{U}(\sigma_\lambda,{\varphi})$ which describe the asymptotic behaviour
of elements of $B_{U}(\sigma_\lambda,{\varphi})$.
For simplicity we assume that the twist ${\varphi}$ is normalized
as in \ref{weih213}.
\subsubsection{}\label{weih244}
We assume that $U\subset P$ defines a cusp of smaller rank.
The case of cusps of full rank will be discussed in
Subsection \ref{maxrank}.
We want to define
a map $\pi^{P_{U}}_*:A_{\{1\}}(\sigma_\lambda,{\varphi})^n\rightarrow
C^\infty(\Omega_P\setminus N_V \{0_P\},V(\sigma_\lambda,{\varphi}))$
by
$$\pi^{P_{U}}_*(f):=\int_{P_U}
\pi^{\sigma_\lambda,{\varphi}}(u) fdu \ .$$
If $x\not\in N_V$, then the integrand $(\pi^{\sigma_\lambda,{\varphi}}(u) f)(xw)$ is well-defined for all $u\in P_{U}$.
We will see below that this integral converges for
${\rm Re }(\lambda)<\rho^U-l_{\varphi}/2$.
Let $S(N_V):=\{x\in N_V|a(xw)=1\}$ (see \ref{weih213} for a definition of $l_{\varphi}$ and Definition \ref{weih215} for $\rho^U$).
\subsubsection{}
Employing the assumption that $N_V$
is invariant under conjugation by $A$ (regularity of the cusp, Definition \ref{t799})
we can consider polar coordinates $S(N_V)\times A\ni (\xi,a)\mapsto \xi^a\in N_V$
of $N_V\setminus \{1\}$.
\begin{lem}\label{210}
There is a measure $d\xi$ on $S(N_V)$ such that the Haar measure $dx$
of $N_V$ is given by $ a^{2\rho_U} dad\xi $.
\end{lem}
{\it Proof.$\:\:\:\:$}
There is a family of measures $d\xi(a)$ on $S(N_V)$
such that for any $f\in C_c(N_V)$ we have
$$\int_{N_V} f(x) dx =\int_A \int_{S(N_V)} f(\xi^a) d\xi(a) a^{2\rho_U} da\ .$$
We compute for any $b\in A$
\begin{eqnarray*}
\int_{N_V} f(x^{b^{-1}}) dx &=& b^{2\rho_U}\int_{N_V} f(x) dx\\
&=& b^{2\rho_U} \int_A \int_{S(N_V)} f(\xi^a) d\xi(a) a^{2\rho_U} da\ ,\\
\int_{N_V} f(x^{b^{-1}}) dx &=&\int_A \int_{S(N_V)} f(\xi^{ab^{-1}}) d\xi(a) a^{2\rho_U} da\\
&=&b^{2\rho_U} \int_A \int_{S(N_V)} f(\xi^a) d\xi(ab) a^{2\rho_U} da\ .
\end{eqnarray*}
We conclude that $d\xi(ab)=d\xi(a)$ for all $b\in A$ and hence $d\xi(a)=d\xi(1)=:d\xi$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}\label{weih221}
Recall the the Definition \ref{weih1422} of $P_U=N_V M_U$.
We introduce the abbreviation $$f_0:= \int_{M_{U}} \pi^{\sigma_\lambda,{\varphi}}(t) f dt\ .$$
Then we can write for $x\not\in N_V$ and using the homogeneity (\ref{weih217})
\begin{eqnarray}
\pi^{P_{U}}_*(f)(xw)&=&\int_A \int_{S(N_V)} {\varphi}(\xi^a)^{-1} f_0( \xi^a xw) a^{2\rho_U} d\xi da\label{hj11}\\&=&\int_A \int_{S(N_V)} {\varphi}(a) {\varphi}(\xi)^{-1} {\varphi}(a)^{-1}f_0( a\xi a^{-1} xw) a^{2\rho_U} d\xi da\nonumber\\
&=&\int_A \int_{S(N_V)} {\varphi}(a) {\varphi}(\xi)^{-1} {\varphi}(a)^{-1}f_0( a\xi x^{a^{-1}}a^{-1} w) a^{2\rho_U} d\xi da\nonumber\\
&=& \int_A\int_{S(N_V)} {\varphi}(a) {\varphi}(\xi)^{-1} f_0(\xi x^{a^{-1}}w) a^{2(\lambda-\rho^U)-n\alpha} d\xi da\ .\label{hj12}
\end{eqnarray}
We decompose the outer integral into the integrals over $A_+$ and $A_-$
and obtain $\pi^{P_{U} }_+(f)$, $\pi^{P_{U} }_- (f)$ such that $\pi^{P_{U}}_*(f)=\pi^{P_{U} }_+(f) +\pi^{P_{U} }_-(f) $.
It is clear from (\ref{hj11}) that $\pi^{P_{U} }_-(f)$ converges for all $\lambda\in{\aaaa_\C^\ast}$.
We conclude from (\ref{hj12}) that $\pi^{P_{U}}_+(f)$ converges for
${\rm Re }(\lambda)<\rho^U+(n\alpha-l_{\varphi})/2$.
In this domain of convergence we compute for $b\in A$
$${\varphi}(b)^{-1}\pi^{P_{U} }_*(f)(x^bw)= b^{2(\lambda-\rho^U)-n\alpha}\pi^{P_{U} }_*(f)(xw)\ .$$
\subsubsection{}
Motivated by this calculation we make the following definition.
\begin{ddd}
We define $A_{P_U}(\sigma_\lambda,{\varphi})^n$ to be the space
of all $P_U$-invariant $f\in C^\infty(\Omega_P\setminus N_V 0_P,V(\sigma_\lambda,{\varphi}))$
satisfying $${\varphi}(a)^{-1}f(x^aw)=a^{2(\lambda-\rho^U)-n\alpha}
f(xw)$$ for all $a\in A$, $x\in N\setminus N_V$.
\end{ddd}
\subsubsection{}
\begin{lem}
The family of spaces $\{A_{P_U}(\sigma_\lambda,{\varphi})^n\}_{\lambda\in{\aaaa_\C^\ast}}$ forms a trivial holomorphic family
of Fr\'echet and Montel spaces.
\end{lem}
{\it Proof.$\:\:\:\:$}
For some $z\in{\mathbb{R}}$ with $z<-\rho_U/\alpha$
we define the function
$$s_1(x):=\int_{P_{U}} a(y.xw)^{z\alpha} dy\ .$$
We further define $s$ as the $\alpha/(z\alpha+\rho_U)$'th power of $s_1$.
Now we consider the function $s$
as a section $s\in C^\infty(\Omega_P\setminus N_V0_P,V(1_{\rho+\alpha}))$
defining $s(xw):= s(x)$.
Multiplication by $s^\mu$ defines
an continuous isomorphism of $A_{P_U}(\sigma_\lambda,{\varphi})^n$
with $A_{P_U}(\sigma_{\lambda+\mu},{\varphi})^n$. We employ these isomorphisms in order
to define a holomorphic trivialization of the bundle.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
For ${\rm Re }(\lambda)<\rho^U+(n\alpha-l_{\varphi})/2$ we have defined above a holomorphic family of continuous maps
$$\pi^{P_{U} }_*:A_{\{1\}}(\sigma_\lambda,{\varphi})^n\rightarrow A_{P_{U}}(\sigma_\lambda,{\varphi})^n\ .$$
\begin{lem}\label{aaww}
The family $\pi^{P_{U} }_*$ extends meromorphically to all of ${\aaaa_\C^\ast}$ with at most first order poles in the set $\rho^U+\frac{n\alpha-l_{\varphi}}{2} +\frac{1}{2}{\mathbb {N}}_0$. The residues are finite-dimensional.
\end{lem}
{\it Proof.$\:\:\:\:$}
Recall the decomposition $\pi^{P_{U} }_*=\pi^{P_{U} }_++\pi^{P_{U} }_-$ introduced in \ref{weih221}. We have already seen that $\pi^{P_{U} }_-$ has a holomorphic continuation.
It suffices to show that $\pi^{P_{U} }_+$ has a meromorphic continuation.
Consider $f\in A_{\{1\}}(\sigma_\lambda,{\varphi})^n$.
Let $$F(x):=\int_{S(N_V)} {\varphi}(\xi)^{-1} f_0(\xi xw) d\xi$$
(see \ref{weih221} for the definition of $f_0$).
This function is smooth in a small neighbourhood
of $1\in N$ and on $N\setminus N_V$. Using the Taylor formula
for each $r\in{\mathbb {N}}_0$ we can write
$F(x^a)=\sum_{q=0}^r F_q(x) a^{q\alpha}+ a^{(r+1)\alpha}R_r(x,a)$,
where $F_q$ is a homogeneous polynomial of degree $q\alpha$ on $N$
and $R_r(x,a)$ is uniformly bounded as $a\to 0$.
We write
$\pi^{P_{U} }_+(f)(xw):=\int_{A_+} {\varphi}(a) F(x^{a^{-1}}) a^{2(\lambda-\rho^U)-n\alpha}da$
and insert the expansion for $F$ in order to obtain
$$\pi^{P_{U} }_+(f)(xw):=J_r^1(f)(xw)+J_r^2(f)(xw)\ ,$$ where
\begin{eqnarray*}
J_r^1(f)(xw)&:=& \sum_{q=0}^r \int_{A_+} {\varphi}(a) a^{2(\lambda-\rho^U)-(q+n)\alpha}da F_q(x)\\
J_r^2(f)(xw)&:= &\int_{A_+} {\varphi}(a)R_r(x,a^{-1}) a^{2(\lambda-\rho^U)-(r+1+n)\alpha}da\ .\end{eqnarray*}
The integral $J_r^2$ converges for ${\rm Re }(\lambda)<\rho^U+ (n+r+1)\alpha/2-l_{\varphi}/2$.
In order to evaluate $J_r^1$ we introduce the operator $B:=(d/da)_{|a=1}{\varphi}(a)\in {\mathrm{End}}(V_{\varphi})$.
In order to define this derivative we embed $A$ into the multiplicative group of ${\mathbb{R}}$.
Since ${\varphi}$ is assumed to be algebraic as a representation of $A$ the eigenvalues of $B$ are integral.
We have
$$J_r^1(f)(xw) = - \sum_{q=0}^r (B+2(\lambda-\rho^U)-(n+q)\alpha)^{-1} F_q(x)\ .$$
In particular we see that $J_r^1$ has a meromorphic continuation
to all of ${\aaaa_\C^\ast}$.
Since we can choose $r$ arbitrarily large we obtain a meromorphic continuation
of $\pi^{P_{U} }_+$ to all of ${\aaaa_\C^\ast}$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Recall the Definition \ref{weih231} of the subspace of polynomials
${R}_{\{1\}}(\sigma_\lambda,{\varphi})^n\subset A_{1}(\sigma_\lambda,{\varphi})^n$.
We consider the following diagram
$$\xymatrix{{R}_{\{1\}}(\sigma_\lambda,{\varphi})^n\ar[d]\ar[r]&A_{1}(\sigma_\lambda,{\varphi})^n\ar[d]^{\pi^{P_{U} }_*}\\?\ar[r]&A_{P_{U}}(\sigma_\lambda,{\varphi})^n}\ .$$
We can complete the diagram using the concept of an image bundle Definition \ref{imagebundle}.
\begin{ddd}\label{weih254}
We define
${R}_{U}(\sigma_\lambda,{\varphi})^n \subset A_{P_{U}}(\sigma_\lambda,{\varphi})^n$
as the image bundle
of the restriction of $\pi^{P_{U} }_*$ to ${R}_{\{1\}}(\sigma_\lambda,{\varphi})^n$.
We define the meromorphic family of maps
$$[\pi^{U}_*]:{R}_{\{1\}}(\sigma_\lambda,{\varphi})^n\rightarrow {R}_{U}(\sigma_\lambda,{\varphi})^n$$
to be induced by $\pi^{P_{U}}_*$. Furthermore, we define
${R}_{U,k}(\sigma_\lambda,{\varphi}):=\bigoplus_{n=0}^k {R}_{U}(\sigma_\lambda,{\varphi})^n$
and let $$[\pi^{U}_*]:{R}_{\{1\},k}(\sigma_\lambda,{\varphi})\rightarrow {R}_{U,k}(\sigma_\lambda,{\varphi})$$
denote the corresponding meromorphic family of maps.
\end{ddd}
Similar to the usage of $AS$ and $L$
the symbol $[\pi^{\Gamma}_*]$ denotes various versions of the push-down
on the level of asymptotic terms. It will be clear from the context wich version is meant.
\subsubsection{}\label{weih500}
Using the second assertion of Lemma \ref{bunle}
we choose once and for all meromorphic families of right-inverses $$[Q]: {R}_{U}(\sigma_\lambda,{\varphi})^n \rightarrow {R}_{\{1\}}(\sigma_\lambda,{\varphi})^n $$
of $[\pi^{U}_*]$.
\subsubsection{}\label{weih253}
Let $\chi\in C^\infty(B_U)$ be a cut-off function such that
$1-\chi$ has compact support, and which vanishes in a neighbourhood
of $U\backslash (P_U 0_P)$. We can assume that $\chi$
is $P_U$-invariant (otherwise we replace it by
the average $xw\mapsto \int_{U\backslash P_U} \chi(u^{-1}xw) du$).
Multiplication by $\chi$ defines
an inclusion $L:{R}_{U,k}(\sigma_\lambda,{\varphi})\rightarrow
C^\infty(B_{U},V_{B_{U}}(\sigma_\lambda,{\varphi}))$.
As in Subsection \ref{trivas} we see that
$L({R}_{U,k}(\sigma_\lambda,{\varphi}))\cap {S}_{U,k}(\sigma_\lambda,{\varphi})=\{0\}$.
\subsubsection{}\label{neuj915}
\begin{ddd}\label{neuj104}
We define the Fr\'echet space
$$B_{U,k}(\sigma_\lambda,{\varphi}):= {S}_{U,k}(\sigma_\lambda,{\varphi})\oplus L({R}_{U,k}(\sigma_\lambda,{\varphi}))\ .$$ Furthermore we
define the Fr\'echet and Montel space $B_{U}(\sigma_\lambda,{\varphi})$ to be the intersection
of the spaces $B_{U,k}(\sigma_\lambda,{\varphi})$ for all $k\in{\mathbb {N}}$.
\end{ddd}
Note that $B_{U,k}(\sigma_\lambda,{\varphi})$ fits into the split exact sequence
\begin{equation}\label{sse0}0\rightarrow {S}_{U,k}(\sigma_\lambda,{\varphi})\rightarrow B_{U,k}(\sigma_\lambda,{\varphi})
\stackrel{AS}{\rightarrow} {R}_{U,k}(\sigma_\lambda,{\varphi})\rightarrow 0\ ,\end{equation}
where $AS$ takes the finite asymptotic expansion.
Since the spaces ${S}_{U,k}(\sigma_\lambda,{\varphi})$ and ${R}_{U,k}(\sigma_\lambda,{\varphi})$ form trivial holomorphic bundles we can employ the split $L$ of the sequence
in order to equip the family of spaces $\{B_{U,k}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$
with the structure of a trivial holomorphic bundle over ${\aaaa_\C^\ast}$.
The space
$B_{U}(\sigma_\lambda,{\varphi})$
fits into the exact sequence
\begin{equation}\label{sse1}0\rightarrow S_{U}(\sigma_\lambda,{\varphi})\rightarrow
B_{U}(\sigma_\lambda,{\varphi})\rightarrow {R}_{U}(\sigma_\lambda,{\varphi})\rightarrow 0\end{equation}
which does not admit any continuous split.
The spaces $B_{U}(\sigma_\lambda,{\varphi})$ form a projective limit
of locally trivial holomorphic bundles in the sense of \ref{weih211}.
\subsubsection{}
We now show compatibility of the spaces
$B_{U}(\sigma_\lambda,{\varphi})$ with twisting.
We assume that $\sigma$ is Weyl invariant as in \ref{weih241}.
Let $(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$ be a finite-dimensional
representation of $G$ as in \ref{weih144}.
Recall the notation $p_{\sigma,\mu}$ and $i_{\sigma,\mu}$ from \ref{weih142}.
\begin{lem}\label{compat2}
\begin{enumerate}
\item For each $n\in{\mathbb {N}}_0$ we have the following commutative diagrams
\begin{eqnarray}&&
\begin{array}{ccc}
R_{\{1\}}(\sigma_\lambda,{\varphi})^n&\stackrel{i_{\sigma,\mu}^n}{\rightarrow}&
R_{\{1\}}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m\\
\downarrow [\pi^{U}_*]&&\downarrow [\pi^{U}_*]\\
R_{\{U\}}(\sigma_\lambda,{\varphi})^n&\stackrel{[i_{\sigma,\mu}^{U}]}{\rightarrow}&
R_{\{U\}}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m
\end{array}\label{uz76}\\&&
\begin{array}{ccc}
R_{\{1\}}(\sigma_\lambda,{\varphi})^n&\stackrel{p_{\sigma,\mu}^n}{\leftarrow}&
R_{\{1\}}(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m\\
\downarrow [\pi^{U}_*]&&\downarrow [\pi^{U}_*]\\
R_{\{U\}}(\sigma_\lambda,{\varphi})^n&\stackrel{[p_{\sigma,\mu}^{U}]}{\leftarrow}&
R_{\{U\}}(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m
\end{array} \ ,\nonumber
\end{eqnarray}
where $m:=n+\mu/\alpha$.
\item
We have holomorphic families of continuous maps
$$i_{\sigma,\mu}^U:B_{U}(\sigma_\lambda,{\varphi})\rightarrow
B_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$$
and
$$p_{\sigma,\mu}^U:B_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow
B_{U}(\sigma_\lambda,{\varphi})\ .$$
\item
$B_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
is a $Z({\mathfrak{g}})$-module.
\item
If $\lambda\not\in I_{\mathfrak{a}}$, then
$i_{\sigma,\mu}^U$ maps $B_{U}(\sigma_\lambda,{\varphi})$
isomorphically onto $$\ker\{Z(\lambda):B_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow B_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})\}\ ,$$
and the restriction of
$p_{\sigma,\mu}^U$ to $$\ker\{\tilde Z(\lambda):B_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow B_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\}$$
is an isomorphism onto $B_{U}(\sigma_\lambda,{\varphi})$
(here $\tilde Z(\lambda)$ is the adjoint of
$\pi^{1_{-\lambda+\mu}\otimes\tilde\pi_{\sigma,\mu}, {\mathrm {id}}_{\tilde{\varphi}}}(\Pi(\lambda))$).
\item
The composition
$$j_{\sigma,\mu}^U:B_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})
\stackrel{1-Z(\lambda)}{\rightarrow} \ker_U(Z(\lambda))\stackrel{(i_{\sigma,\mu}^U)^{-1}}{\rightarrow}
B_{U}(\sigma_\lambda,{\varphi})$$
which is initially defined for $\lambda\not\in I_{\mathfrak{a}}$
extends to a meromorphic family of continuous maps.
Similarly, the composition
$$q_{\sigma,\mu}^U:B_{U}(\sigma_\lambda,{\varphi})\stackrel{(p_{\sigma,\mu}^U)^{-1}}{\rightarrow} \ker(\tilde Z(\lambda)) \rightarrow B_U(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})$$
extends to a meromorphic family of continuous maps.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
We consider the first diagram of 1. and leave the second to the reader
since the argument is similar.
First one checks that the inclusion
$i_{\sigma,\mu}$ induces an inclusion
$$i_{\sigma,\mu}^{P_U,n}:{\cal A}_{P_U}(\sigma_\lambda,{\varphi})^n\rightarrow
{\cal A}_{P_U}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^{m}$$
for all $n\in{\mathbb {N}}_0$, $m:=n+\mu/\alpha$. Now we obtain diagram
(\ref{uz76}),
where the map $[i_{\sigma,\mu}^{U}]$ is the restriction
of $i_{\sigma,\mu}^{P_U,n}$ which is well-defined by the
naturality of the image bundle construction Lemma \ref{bunle}.
\subsubsection{}
Assertion 2. follows from Lemma \ref{compat1}, 2. and 1.
Indeed, $i^{U}_{\sigma,\mu}$ maps
$B_{U,k}(\sigma_\lambda,{\varphi})$ to
$$\{i^U_{\sigma,\mu}\}(\{S_{U,k}(\sigma_\lambda,{\varphi})\})+L([i_{\sigma,\mu}^{U}](R_{U,k}(\sigma_\lambda,{\varphi})))
\subset B_{U,k^\prime}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi}) \ ,$$
where $k^\prime=k+\mu/\alpha$. The argument for the second
assertion of 2. is similar.
\subsubsection{}
In order to prove 3. we first show that the spaces
$R_{U}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m$ are a ${\mathcal{Z}}({\mathfrak{g}})$-modules.
Note that the representations $\pi^{1_{\lambda+\mu}\otimes\pi_{\sigma,\mu},{\mathrm {id}}_{\varphi}}$
of ${\mathcal{Z}}({\mathfrak{g}})$ and $\pi^{1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi}}$
of $AP_U$ commute. Since we employ the latter representation in order to define homogeneities
we see that the spaces $A_{P_U}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m$
are ${\mathcal{Z}}({\mathfrak{g}})$-modules. The space $R_{\{1\}}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
is clearly a $Z({\mathfrak{g}})$-module since it is the quotient
of $C^\infty(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi}))$
by the submodule $S_{\{1\}}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
(Lemma \ref{compat1}, 3.). Thus the subspaces
$R_{\{1\}}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m$ are
$Z({\mathfrak{g}})$-modules. We obtain the $Z({\mathfrak{g}})$-module structure
on $R_{U}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m$
from Lemma \ref{bunle} since $\pi^{P_U}_*$ is ${\mathcal{Z}}({\mathfrak{g}})$-equivariant.
Assertion 3. now follows from Lemma \ref{compat1}, 3. and the fact that for any $A\in {\mathcal{Z}}({\mathfrak{g}})$ the commutator
$[\pi^{1_{\lambda+\mu}\otimes\pi_{\sigma,\mu},{\mathrm {id}}_{\varphi}}(A),\chi]$
is a differential operator with compactly supported coefficients on
$B_U$.
\subsubsection{}
We now prove the first assertion of 4. and leave the second to the reader.
The exact sequence of bundles (\ref{hj112}) induces an exact sequence
$$0\rightarrow R_U(\sigma_\lambda,{\varphi})^n\stackrel{[i^U_{\sigma,\mu}]}{\rightarrow}
R_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m
\stackrel{p}{\rightarrow} R_U(w,{\varphi})^k\ ,$$
where $m=n+\mu/\alpha$ and we employ an appropriate definition of
homogeneity for the last space. If $\lambda\not\in I_{\mathfrak{a}}$,
then we have $\ker(Z(\lambda):R_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m
\rightarrow R_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})^m)=\ker(p)$.
For these $\lambda$ the map $[i^U_{\sigma,\mu}]$ identifies
$R_U(\sigma_\lambda,{\varphi})^n$ with the kernel of $Z(\lambda)$.
If $f\in B_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
satisfies $Z(\lambda)(f)=0$, then $Z(\lambda)AS(f)=0$.
From what was shown above and (\ref{sse1}) it follows
that there exists $g\in B_U(\sigma_\lambda,{\varphi})$
such that $[i^U_{\sigma,\mu}]\circ AS(g) = AS(f)$.
Then $f-i^U_{\sigma,\mu}(g)\in S_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$,
and by Lemma \ref{compat1}, 4. there exists $h\in S_U(\sigma_\lambda,{\varphi})$
such that $\{i^U_{\sigma,\mu}\}(h)=f-i^U_{\sigma,\mu}(g)$.
Thus $f= i^U_{\sigma,\mu}(h+g)$, and this finishes the proof of 4.
\subsubsection{}
We now prove the first assertion of 5. and leave the second to the reader.
We employ the notation introduced in the proof of Lemma \ref{compat1}, 5.
For each $n\in{\mathbb {N}}_0$ there is $m\in{\mathbb {N}}_0$ such that
$J:S_{U,m}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi}) \rightarrow
S_{U,n}(\sigma_\lambda,{\varphi})$ is a holomorphic family of maps.
We define the meromorphic family of continuous maps
$\tilde j^U_{\sigma,\mu}:B_{U,m}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi}) \rightarrow B_{U,n}(\sigma_\lambda,{\varphi})$ by
$$\tilde j^U_{\sigma,\mu}(f):=
L\circ [i^U_{\sigma,\mu}]^{-1}\circ AS\circ (1-Z(\lambda))(f) +
\{j^U_{\sigma,\mu}\}(f- L\circ AS \circ (1-Z(\lambda))(f))\ .$$
Then $\tilde j^U_{\sigma,\mu}\circ i^U_{\sigma,\mu}(f)=f$
for any $f\in B_{U}(\sigma_\lambda,{\varphi})$.
Since $\tilde j^U_{\sigma,\mu}$ vanishes on $\ker(1-Z(\lambda))$
we conclude that
the restriction of $\tilde j^U_{\sigma,\mu}$ to
$B_{U}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
coincides with $j^U_{\sigma,\mu}$ for $\lambda\not\in I_{\mathfrak{a}}$.
This proves 5. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsection{The push-down for Schwartz spaces}\label{psuscg}
\subsubsection{}
In this subsection we show that the push-down induces a map
between Schwartz spaces
$$\{\pi^U_*\}:{S}_{\{1\},k}(1_\lambda,{\varphi})\rightarrow {S}_{U,k}(1_\lambda,{\varphi}) \ .$$
Note that we only consider the spherical case $\sigma=1$.
\begin{lem}\label{schwpush}
\begin{enumerate}
\item The push-down $\{\pi^U_*\}$ converges for
${\rm Re }(\lambda)<\rho^U+(k\alpha-l_{\varphi})/2$.
\item For $k_1<k$ it induces a holomorphic
family of maps
$\{\pi^U_*\}:{S}_{\{1\},k}(1_\lambda,{\varphi})\rightarrow {S}_{U,k_1}(1_\lambda,{\varphi})$.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
Let $f\in C_c^\infty(B_{\{1\}},V_{B_{\{1\}}}(1_\lambda,{\varphi}))$.
Then $\{\pi^{U}_*\}(f)=\sum_{u\in U}\pi^{1_\lambda,{\varphi}}(u)f$
converges and defines an element of
$C_c^\infty(B_{U},V_{B_{U}}(1_\lambda,{\varphi}))$.
\subsubsection{}
Let $h\in{\mathbb {N}}_0$, $W\subset N\setminus N_V$ be compact, $|.|$
be a ${\varphi}(M_{U})$-invariant norm on $V_{\varphi}$, and
$D\in {\mathcal{U}}({\mathfrak{n}})^{\le d \alpha}$ (see \ref{weih243} for notation). Then we consider the seminorm
$$q_{W,D,h}(f):=\sup_{x\in W}\sup_{a\in A_+} a^{h\alpha-2(\lambda-\rho^U)}|{\varphi}(a)^{-1}(f)(x^aDw)|\ .$$
The first assertion of the Lemma will easily follow from estimates of the form
$$q_{W,D,k}(\{\pi^{U}_*\}f)<C\|f\|_{k,d}\ ,$$
and the second assertion will follow from
the convergence
of sums of the form
$$\sum_{u\in U} q_{W,D,k_1}(\pi^{1_\lambda,{\varphi}}(u) f)<C\|f\|_{k,d} \ .$$
\subsubsection{}
If $u\not=1$, then we can write
$u^{-1}=m_u\xi_u^{b_u}$ for suitable $\xi_u\in S(N_V)$, $b_u\in A$, and $m_u\in M_{U}$ (see \ref{weih244} for the notation $S(N_V)$).
There is a constant $C\in R$ such that for all $f\in C_c^\infty(B_{\{1\}},V_{B_{\{1\}}}(1_\lambda,{\varphi}))$,
$x\in W$, $a\in A$, $1\not=u\in U$ with
$b_u\ge a$ we have the estimate
\begin{eqnarray*}
|{\varphi}(a)^{-1}\pi^{1_\lambda,{\varphi}}(u) f(x^aDw)|&=&
|{\varphi}(a)^{-1}{\varphi}(\xi_u^{b_u})^{-1}
f(m_u\xi_u^{b_u}x^aDw)|\\
&=&
|{\varphi}(a^{-1}b_u){\varphi}(\xi_u)^{-1} {\varphi}(b_u^{-1})f(m_u(\xi_ux^{ab_u^{-1}})^{b_u}Dw)|\\
&\le&
C\|f\|_{d,k} |{\varphi}(a^{-1}b_u)| b_u^{2(\lambda-\rho)-k\alpha}
\end{eqnarray*}
and for $b_u\le a$
\begin{eqnarray*}
|{\varphi}(a)^{-1}\pi^{1_\lambda,{\varphi}}(u) f(x^aDw)|&=&
|{\varphi}(a)^{-1} {\varphi}(\xi_u^{b_u})^{-1} f(m_u\xi_u^{b_u}x^aDw)|\\
&=&
|{\varphi}(\xi_u^{a^{-1}b_u})^{-1} {\varphi}(a)^{-1}f(m_u(\xi^{b_ua^{-1}}x)^{a}Dw)|\\
&\le&
C\|f\|_{d,k} a^{2(\lambda-\rho)-k\alpha}\ .
\end{eqnarray*}
\subsubsection{}
If ${\rm Re }(\lambda)<\rho^U+(k\alpha-l_{\varphi})/2$, then
summing up these estimates over $U$ and estimating the sum
by an integral over $P_{U}$ we obtain $C_1\in{\mathbb{R}}$ such that
all $x\in W$, $a\in A^+$, $f\in C_c^\infty(B_{\{1\}},V_{B_{\{1\}}}(1_\lambda,{\varphi}))$
$$|{\varphi}(a)^{-1}\{\pi^{U^0}_*\}(f)(x^aDw)|< C_1 a^{2(\lambda-\rho^U)-k\alpha}\ .$$
This implies $$q_{W,D,k}(\{\pi^{U}_*\}f)<C_1\|f\|_{k,d}\ .$$
\subsubsection{}
Multiplying the two inequalities above by $a^{k_1\alpha-2(\lambda-\rho^U)}$
and taking the supremum over $x$ and $a$ we obtain
$q_{W,D,k_1}(\pi^{1_\lambda,{\varphi}}(u) f)<b_u^{-2\rho_U-(k-k_1)\alpha}$.
Summing this over $U$ and estimating the sum over $U$
by the integral over $P_{U}$ we obtain
$$\sum_{u\in U} q_{W,D,k_1}(\pi^{1_\lambda,{\varphi}}(u) f)<C\|f\|_{k,d}\ .$$
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Recall (Definition \ref{weih245}) that ${S}_{\{1\}}(1_\lambda,{\varphi})$ is the intersection of the spaces
${S}_{\{1\},k}(1_\lambda,{\varphi})$ for $k\in {\mathbb {N}}_0$. Hence Lemma \ref{schwpush}
implies the following corollary.
\begin{kor}\label{tzu}
We have a holomorphic family of maps
$$\{\pi^U_*\}:{S}_{\{1\}}(1_\lambda,{\varphi})\rightarrow {S}_U(1_\lambda,{\varphi})$$
defined on all of ${\aaaa_\C^\ast}$.
\end{kor}
\subsubsection{}
We now turn to the problem of constructing a right-inverse of the push-down for Schwartz spaces.
Its adjoint plays an important role in the construction of the restriction map. Below we encounter the effect of a loss of regularity ($-2\rho_U$). This is one of the places where the presence of cusps makes the theory
much more complicated in comparison with the convex cocompact case.
\subsubsection{}
\begin{lem} \label{schwspl}
For all $k_1\in{\mathbb {N}}_0$ satisfying $k_1 \alpha<k\alpha-2\rho_U$
there exists a holomorphic family of maps
$$\{Q\}:{S}_{U,k}(1_\lambda,{\varphi})\rightarrow {S}_{\{1\},k_1}(1_\lambda,{\varphi})$$
such that $\{\pi^U_*\}\circ \{Q\}$ is the natural inclusion
${S}_{\{1\},k}(1_\lambda,{\varphi})\hookrightarrow {S}_{\{1\},k_1}(1_\lambda,{\varphi})$.
\end{lem}
{\it Proof.$\:\:\:\:$}
We define a holomorphic family $\{Q\}$ of right-inverses of $\{\pi^U_*\}$.
We employ a cut-off function $\chi^U\in C^\infty(\Omega_P)$
with $\sum_{u\in U} u^*\chi^U=1$, and such that
$\chi^U(xD)$ is bounded for each $D\in{\mathcal{U}}({\mathfrak{n}})$. Here we have identified
$\Omega_P$ with $N$. Then $\chi^U$ considered as a function on $N$ can be derived with respect
to the left invariant differential operator $D$.
To see that such a function exists we equip $\Omega_P\cong N$ with a $MN$-invariant Riemannian metric
such that $U$ acts isometrically.
Note that
the Riemannian manifold $U\backslash N$ admits a lower bound of the
injectivity radius.
For $f\in {S}_{U,k}(1_\lambda,{\varphi})$
we define $\{Q\}(f)$ to be the lift of $f$ to $\Omega_P$ multiplied by
$\chi^U$. It is then easy to see that
$\{Q\}: {S}_{U,k}(1_\lambda,{\varphi}) \rightarrow {S}_{\{1\},k_1}(1_\lambda,{\varphi})$ for all $k_1\in{\mathbb {N}}_0$ satisfying $k_1 \alpha<k\alpha-2\rho_U$.
Therefore we obtain a holomorphic family of maps
$\{Q\}:{S}_{U}(1_\lambda,{\varphi})\rightarrow {S}_{\{1\}}(1_\lambda,{\varphi})$.
By definition $\{\pi^U_*\}\circ \{Q\}={\mathrm {id}}$.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsection{Push down for cusps of smaller rank}\label{weih261}
\subsubsection{}
We assume that the cusp associated to $U\subset P$ does not
have full rank.
In order to construct the push-down it remains to extend
$\{\pi^U_*\}$ (constructed in Lemma \ref{schwpush}) for each $k\in{\mathbb {N}}$ from ${S}_{\{1\},k}(1_\lambda,{\varphi})$ to
$B_{\{1\},k}(1_\lambda,{\varphi})$ such that the following diagram
is commutative:
$$
\begin{array}{ccccccccc}
0&\rightarrow&{S}_{\{1\},k}(1_\lambda,{\varphi})&\rightarrow& B_{\{1\},k}(1_\lambda,{\varphi})&\stackrel{AS}{\rightarrow}& {R}_{\{1\},k}(1_\lambda,{\varphi})&\rightarrow &0\\
& &\downarrow \{\pi^{U}_*\} & & \downarrow \pi^{U}_* & & \downarrow [\pi^{U}_*] &&\\
0&\rightarrow&{S}_{U,k}(1_\lambda,{\varphi})&\rightarrow& B_{U,k}(1_\lambda,{\varphi})& \stackrel{AS}{\rightarrow}&{R}_{U,k}(1_\lambda,{\varphi})&\rightarrow &0\ .\end{array}
$$
The main technical result in this direction is the following Proposition \ref{mainpure}.
\subsubsection{}
Note that ${R}_{\{1\}}(1_\lambda,{\varphi})^n$ is a finite-dimensional
representation of the torus $M_{U^0}$ (see \ref{weih251} for the explanation of $M_{U^0}$). Let ${\mathcal{X}}(M_{U^0})$ denote the
set of characters of $M_{U^0}$.
For any $\theta\in {\mathcal{X}}(M_{U^0})$
let ${R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta$ be the subspace
of all $f\in {R}_{\{1\}}(1_\lambda,{\varphi})^n$ such that
$$\pi^{1_\lambda,{\varphi}}(t)f=\theta(t)f\ .$$
Then we have a finite decomposition $${R}_{\{1\}}(1_\lambda,{\varphi})^n=
\bigoplus_{\theta\in{\mathcal{X}}(M_{U^0}) }
{R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta\ .$$
\subsubsection{}
Recall the construction of $L$ from \ref{weih253}.
Furthermore see Definition \ref{weih254} for $[\pi^{U^0}_*]$, Definition \ref{weih215} for $\rho^U$
and \ref{weih213} for $l_{\varphi}$.
\begin{prop}\label{mainpure}
The composition
$$\pi^{U^0}_*\circ L:{R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta\rightarrow
B_{U^0}(1_\lambda,{\varphi})$$
converges for ${\rm Re }(\lambda)<\rho^U+(n\alpha-l_{\varphi})/2$
and has a meromorphic continuation to all of ${\aaaa_\C^\ast}$
such that $AS\circ \pi^{U^0}_*\circ L = [\pi^{U^0}_*]$.
\end{prop}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
We decompose the push-down
$\pi^{U^0}_*$ into two intermediate maps. The first map
is the push-down
$\pi^{[U^0,U^0]}_*\circ L :
{R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta\rightarrow B_{[U^0,U^0]}(1_\lambda,{\varphi})$
with respect to the commutator group
$[U^0,U^0]$, and the second is a relative push-down
$\pi^{U^0/[U^0,U^0]}$.
Note that $M_{U^0}$ centralizes $[U^0,U^0]$ (see \ref{weih131})
and therefore acts on the sequence
$$0\rightarrow {S}_{[U^0,U^0]}(1_\lambda,{\varphi})\rightarrow B_{[U^0,U^0]}(1_\lambda,{\varphi})\rightarrow
{R}_{[U^0,U^0]}(1_\lambda,{\varphi})\rightarrow 0\ .$$
The representation of $M_{U^0}$ on ${R}_{[U^0,U^0]}(1_\lambda,{\varphi})$
is compatible with the decomposition \linebreak[4]
$\bigoplus_{n=0}^\infty {R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n$.
\subsubsection{}\label{weih257}
Let ${R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta$ be the subspace of all
$f\in {R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n$ satisfying $$\pi^{1_\lambda,{\varphi}}(t)f=\theta(t)f\ .$$
Then we have a further finite decomposition
$${R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n=
\bigoplus_{\theta\in{\mathcal{X}}(M_{U^0}) }
{R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta\ .$$
\subsubsection{}
We will first show
\begin{lem} \label{lpro1}
The push-down over $[U^0,U^0]$ defines a meromorphic
family of maps
$$\pi^{[U^0,U^0]}:{R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta\rightarrow
{S}_{[U^0,U^0]}(1_\lambda,{\varphi}) \oplus L({R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta)$$
such that $$AS\circ \pi^{[U^0,U^0]}\circ L=[\pi^{[U^0,U^0]}]\ .$$
\end{lem}
It is clear that the average $\{\pi_*^{U^0/[U^0,U^0]}\}: {S}_{[U^0,U^0]}(1_\lambda,{\varphi})\rightarrow
{S}_{U^0}(1_\lambda,{\varphi})$ over $U^0/[U^0,U^0]$
converges and depends holomorphically
on $\lambda$. A simple way to see this formally is to
write
$$\{\pi_*^{U^0/[U^0,U^0]}\}=\{\pi^{U^0}_*\}\circ\{Q\}\circ \pi_*^{[U^0,U^0]}$$
using the split $\{Q\}$ constructed in Lemma \ref{schwspl}
and Corollary \ref{tzu}.
Proposition \ref{mainpure} now immediately follows from
\begin{lem}\label{lpro2}
The composition
$$\pi^{U^0/[U^0,U^0]}_*\circ L: {R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta
\rightarrow B_{U^0}(1_\lambda,{\varphi})$$
defines a meromorphic family of maps
such that $$AS\circ \pi^{U^0/[U^0,U^0]}_* \circ L \circ [\pi_*^{[U^0,U^0]}]=
[\pi^{U^0}_*]\ .$$
\end{lem}
\subsubsection{}
We now start with the proof of Lemma \ref{lpro1}.
Since $M_{U^0}$ is abelian the homomorphism $m:N_V\rightarrow M_{U^0}$ vanishes
on $[N_V,N_V]$. Therefore
$[U^0,U^0]$ is a discrete subgroup of the center of $N$.
If $f\in {R}_{\{1\}}(1_\lambda,{\varphi})^n$, then for all $d\in{\mathbb {N}}_0$ we have
$\|L(f)\|_{n,d}<\infty$.
Therefore the same arguments as in the proof of Lemma \ref{schwpush} show that the push-down
$\pi^{[U^0,U^0]}_*\circ L$ converges for ${\rm Re }(\lambda)<\rho^{{[U^0,U^0]}}+(n\alpha-l_{\varphi})/2$.
\subsubsection{}\label{weih256}
We consider the abelian group $Z:=P_{{[U^0,U^0]}}=[N_V,N_V]$.
If we are given an
unitary character $\vartheta\in {\mathcal{X}}(Z)$,
then we consider the push-down $\pi^{Z,\vartheta}_*$,
which associates to $f\in C^\infty(\partial X,V(1_\lambda,{\varphi}))$
the average
$$\pi^{Z,\vartheta}_*(f):=\int_{Z} \vartheta(z)^{-1} \pi^{1_\lambda,{\varphi}}(z)(f_{|\Omega}) dz$$
provided the integral converges.
\subsubsection{}
Observe that $[U^0,U^0]$
is a cocompact discrete subgroup of $Z$.
We normalize the Haar measure of $Z$ such that ${\rm vol}({[U^0,U^0]}\backslash Z)=1$.
Let ${\mathcal{X}}(Z,{[U^0,U^0]})$ denote the set of all unitary characters of
$Z$ which are trivial on ${[U^0,U^0]}$. We identify ${\mathcal{X}}(Z)$
with $\imath \mathfrak{z}^*$ and ${\mathcal{X}}(Z,{[U^0,U^0]})$ with a lattice
in $\imath \mathfrak{z}^*$, where $\mathfrak{z}$ denotes the Lie algebra of $Z$.
Our approach is based on the
Poisson summation formula which states that
$$\pi^{[U^0,U^0]}_*(f)=\sum_{\vartheta\in{\mathcal{X}}(Z,{[U^0,U^0]})}
\pi^{Z,\vartheta}_*(f)$$ in the domain of convergence.
\subsubsection{}
We choose an euclidean structure on $\mathfrak{z}$.
This structure induces norms on $\mathfrak{z}$ and $\mathfrak{z}^*$. Furthermore we use
the euclidean structure in order to define a Laplace operator
$\Delta\in {\mathcal{U}}(\mathfrak{z})$ which is normalized such that if
$\vartheta\in{\mathcal{X}}(Z)$, then $\vartheta(\Delta^l x)=|\vartheta|^{2l}
\vartheta(x)$.
Let $COP:{\mathcal{U}}(\mathfrak{z})\rightarrow {\mathcal{U}}(\mathfrak{z})\otimes {\mathcal{U}}(\mathfrak{z})$ denote
the coproduct on ${\mathcal{U}}(\mathfrak{z})$. There are $A_{l,j},B_{l,j},C_{l,j}\in{\mathcal{U}}(\mathfrak{z})$
such that
$$(COP\otimes 1)\circ COP(\Delta^l)=\sum_{j} A_{l,j}\otimes B_{l,j}\otimes C_{l,j}\ .$$
Let $D\mapsto \tilde D$ be the canonical antiautomorphism of ${\mathcal{U}}(\mathfrak{z})$.
\subsubsection{}
Fix $l\in{\mathbb {N}}$ and let $f\in {R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta$.
Recall that $L(f)(xw)=\chi(a(xw)) f(xw)$, where $\chi\in C^\infty(A)$
was some cut-off function which vanishes on a neighbourhood of $A_-$
and is equal to one near $\infty$.
By $S(Z)$ we denote the unit-sphere in $Z$.
For $\xi\in S(Z)$ and $a\in A$ we compute
\begin{eqnarray}
&&\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f)(.w)\nonumber\\
&=&\pi^{1_\lambda,{\varphi}}(a)\pi^{1_\lambda,{\varphi}}((\Delta^l)^{a^{-1}}\xi) \pi^{1_\lambda,{\varphi}}(a^{-1})\chi (a(.w))f(.w)\nonumber\\
&=&a^{\rho-\lambda-4l\alpha}\pi^{1_\lambda,{\varphi}}(a)\pi^{1_\lambda,{\varphi}}(\Delta^l\xi) {\varphi}(a^{-1}) \chi(a(a.a^{-1}w)) f(a. a^{-1}w) \nonumber \\
&=&a^{\lambda-\rho-(n+4l)\alpha}\pi^{1_\lambda,{\varphi}}(a)\pi^{1_\lambda,{\varphi}}(\Delta^l) {\varphi}(\xi) \chi(a^2a(\xi^{-1}.w)) f(\xi^{-1}.w)\nonumber\\
&=&a^{\lambda-\rho-(n+4l)\alpha}\sum_{j}
\pi^{1_\lambda,{\varphi}}(a) {\varphi}(A_{l,j}\xi) \chi(a^2 a(\xi^{-1}\tilde B_{l,j}.w)) f(\xi^{-1}\tilde C_{l,j}.w)\nonumber\\
&=&
a^{2(\lambda-\rho)-(n+4l)\alpha}\sum_{j}
{\varphi}(a) {\varphi}(A_{l,j}\xi) \chi(a^2 a(\xi^{-1}\tilde B_{l,j}a^{-1}.aw)) f(\xi^{-1}\tilde C_{l,j}a^{-1}.aw)\label{zzz}\ .
\end{eqnarray}
\subsubsection{}
If $W\subset N$ is any compact subset and $D\in {\mathcal{U}}({\mathfrak{n}})$, then there
is a constant $C\in{\mathbb{R}}$ such that
$$|\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f)(yDw)|\le C a^{2(\lambda-\rho)-(n+4l)\alpha+l_{\varphi}}$$
for all $\xi\in S(Z)$ and $a\in A_+$.
If ${\rm Re }(\lambda)$ is sufficiently small, then by partial integration for
$\vartheta\in {\mathcal{X}}(Z,{[U^0,U^0]})$, $\vartheta\not=0$
\begin{eqnarray*}
\pi^{Z,\vartheta}_*(L(f))&=&\int_{Z}
\vartheta(x)^{-1}\pi^{1_\lambda,{\varphi}}(x) L(f) dx\\
&=&\frac{1}{|\vartheta|^{2l}}\int_{Z} \vartheta(\Delta^lx)^{-1} \pi^{1_\lambda,{\varphi}}(x) L(f) dx\\
&=&\frac{1}{|\vartheta|^{2l}}\int_{Z} \vartheta(x)^{-1}\pi^{1_\lambda,{\varphi}}(\Delta^l x) L(f) dx\\
&=&\frac{1}{|\vartheta|^{2l}}\int_{S(Z)} \int_A
\vartheta(\xi^a)^{-1}\pi^{1_\lambda,{\varphi}}(\Delta^l \xi^a) L(f)a^{2\rho_{{[U^0,U^0]}}} da d\xi\ .
\end{eqnarray*}
By the estimate above
the integral converges locally uniformly on
$\{{\rm Re }(\lambda)<\rho^{{[U^0,U^0]}}+(n+4l)\alpha/2 -l_{\varphi}/2\}$,
and for any compact subset of this region,
$D\in{\mathcal{U}}({\mathfrak{n}})$, and compact subset $W\subset N$
there is a constant $C_1$ such that
$$|\pi^{Z,\vartheta}_*(L(f))(yDw)|\le \frac{C_1}{ |\vartheta|^{2l}}$$
for all $y\in W$.
\subsubsection{}
Choosing $2l>\dim(Z)$ we see that the sum
$$\sum_{0\not=\vartheta\in{\mathcal{X}}(Z,{[U^0,U^0]})}
\pi^{Z,\vartheta}_*(L(f))$$ converges in $C^\infty(\Omega_P,V(1_\lambda,{\varphi}))$.
\subsubsection{}
Refining the estimates above we now show that this
sum in fact converges in the space of rapidly decreasing functions.
Let $W_1\subset N\setminus Z$ be a compact subset and $D\in{\mathcal{U}}({\mathfrak{n}})$.
It immediately follows from (\ref{zzz}), that there is a constant $C_2$
such that for all $\xi\in S(Z)$, $a,b\in A_+$, with $a \ge b$ and $y\in W_1$
$$|{\varphi}(b)^{-1}(\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f))(y^bDw)|\le
C_2 a^{2(\lambda-\rho)-(n+4l)\alpha}| {\varphi}(ab^{-1})| \ .$$
Let $COP(D)=\sum_{h} D_h^l\otimes D_h^r$
denote the coproduct of $D$, where here
$COP:{\mathcal{U}}({\mathfrak{n}})\rightarrow {\mathcal{U}}({\mathfrak{n}})\otimes{\mathcal{U}}({\mathfrak{n}})$.
We find a constant $C_3\in {\mathbb{R}}$ such that
for all $\xi\in S(Z)$, $a,b\in A_+$ with $a \le b$ and $y\in W_1$
\begin{eqnarray*}
&&|{\varphi}(b)^{-1}(\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f))(y^bDw)|\\
&=&
a^{2(\lambda-\rho)-(n+4l)\alpha}|\sum_{j,h}
{\varphi}(ab^{-1}) {\varphi}(A_{l,j}\xi) \chi(a^2 a(\xi^{-1}\tilde B_{l,j}a^{-1}y^bD^l_haw)) f(\xi^{-1}\tilde C_{l,j}a^{-1}y^bD^r_haw)|\\
&=&
a^{-4l\alpha}
b^{2(\lambda-\rho)-n\alpha}
|\sum_{j,h} {\varphi}((A_{l,j}\xi)^{ab^{-1}}) \chi(b^2 a((\xi^{-1}\tilde B_{l,j})^{b^{-1}a}y(D_h^l)^{b^{-1}}w)) f((\xi^{-1}\tilde C_{l,j})^{b^{-1}a}y(D_h^r)^{b^{-1}}w)|\\
&\le& C_3 b^{2(\lambda-\rho)-(n+4l)\alpha}\ .
\end{eqnarray*}
Using this estimate we see that
$${\varphi}(b)^{-1}\sum_{0\not=\vartheta\in{\mathcal{X}}(Z,{[U^0,U^0]})}
\pi^{Z,\vartheta}_*(L(f))(y^bDw)$$
can be estimated by $C b^{2(\lambda-\rho^{{[U^0,U^0]}})-(n+4l)\alpha}$,
where $C$ can be choosen uniformly for $y\in W_1$ and $\lambda$ in compact subsets
of $\{{\rm Re }(\lambda)<\rho^{{[U^0,U^0]}}+(n+4l)\alpha/2 -l_{\varphi}/2\}$.
Since we can choose $l$ arbitrary large we obtain
a holomorphic continuation of the sum above to all
of ${\aaaa_\C^\ast}$. Moreover we see that this sum
is rapidly decreasing with respect to $b$.
\subsubsection{}
We now consider $\pi^{Z,0}_*(L(f))$.
If ${\rm Re }(\lambda)$ is sufficiently small then we can write
$$\pi^{Z,0}_*(L(f))=\int_A \int_{S(Z)} \pi^{1_\lambda,{\varphi}}(\xi^a) L(f) d\xi a^{2\rho_{[U^0,U^0]}} da\ .$$
We again employ the relation
\begin{eqnarray*}
&&\pi^{1_\lambda,{\varphi}}(\xi^a) L(f)(yw)\\
&=&
a^{2(\lambda-\rho)-n\alpha}
{\varphi}(a) {\varphi}(\xi) \chi(a^2 a(\xi^{-1}a^{-1}yaw)) f(\xi^{-1}a^{-1}yaw)\ .
\end{eqnarray*}
Note that
$$F(a,y):= \int_{S(Z)}{\varphi}(\xi) \chi(a^2 a(\xi^{-1}yw)) f(\xi^{-1}yw) d\xi$$
is a smooth function of $y$ near $y=0$ which is independent
of $a$ for large $a$.
Let
$$F(a,y^{a^{-1}})=\sum_{q=0}^r F_q(y)a^{-q\alpha} + a^{-(r+1)\alpha} R_r(a,y)$$
be the asymptotic expansion for large $a\in A$ obtained from the Taylor
series of $F(a,y)$ at $y=0$. The remainder $R_r(a,y)$ remains bounded
as $a\to\infty$. We write
$$\pi^{Z,1}_*(L(f))(yw)=I_+(y)+I_-(y)\ ,$$ where
\begin{eqnarray*}
I_+(y)&:=&\int_{A_+} a^{2(\lambda-\rho^{[U^0,U^0]})-n\alpha} {\varphi}(a) F(a,y^{a^{-1}}) da\\
I_-(y)&:=&\int_{A_-}\int_{S(Z)} \pi^{1_\lambda,{\varphi}}(\xi^a) L(f)(yw) d\xi a^{2\rho_{[U^0,U^0]}} da
\ .\end{eqnarray*}
The integral $I_-$ converges for all $\lambda\in{\aaaa_\C^\ast}$
and defines a holomorphic family of smooth functions.
We write $I_+(y):=J_r^1(y)+J_r^2(y)$, where
\begin{eqnarray*}
J_r^1(y)&:=& \int_{A_+} a^{2(\lambda-\rho^{[U^0,U^0]})-n\alpha} {\varphi}(a) \sum_{q=0}^r F_q(y)a^{-q\alpha} da\\
J_r^2(y)&:=& \int_{A_+} a^{2(\lambda-\rho^{[U^0,U^0]})-n\alpha} {\varphi}(a)
a^{-(r+1)\alpha} R_r(a,y) da\ .
\end{eqnarray*}
The integral $J_r^2$
converges for ${\rm Re }(\lambda)<\rho^{[U^0,U^0]}+(n+r+1)\alpha/2-l_{\varphi}/2$
and defines a smooth function in $y$. The integral $J_r^1$ can be
evaluated:
$$J_r^1(y)=-\sum_{q=0}^r (B+2(\lambda-\rho^{[U^0,U^0]})-(n+q)\alpha)^{-1} F_q(y)\ ,$$
where $B:=\frac{d}{da}_{|a=1}{\varphi}(a)\in {\mathrm{End}}(V_{\varphi})$ (compare with the proof of Lemma \ref{aaww}).
It obviously defines a meromorphic family of smooth functions.
Since we can choose $r$ arbitrary large we obtain a meromorphic continuation of
$\pi^{Z,0}_*(L(f))$ to all of ${\aaaa_\C^\ast}$. Note that if $y\in N\setminus Z$ and
$b\in A$
is sufficiently large, then we have $\pi^{Z,0}_*(L(f))(y^bw)=\pi_*^{P_{[U^0,U^0]}}(f)(y^bw)$,
where
$\pi_*^{P_{[U^0,U^0]}}$
was discussed in Subsection \ref{asz}.
We conclude that
$\pi^{Z,0}_*(L(f))\in S_{{[U^0,U^0]}}(1_\lambda,{\varphi})\oplus L({R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n)$.
We have shown that
$$\pi^{[U^0,U^0]}_*\circ L:{R}_{\{1\}}(1_\lambda,{\varphi})^n\rightarrow
S_{{[U^0,U^0]}}(1_\lambda,{\varphi})\oplus L({R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta)\subset
B_{{[U^0,U^0]}}(1_\lambda,{\varphi})$$
has a meromorphic continuation to all of ${\aaaa_\C^\ast}$ such that
$AS\circ \pi^{[U^0,U^0]}_*\circ L = [\pi^{[U^0,U^0]}_*]$.
This finishes the proof of the lemma.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{} \label{weih258}
We now prove Lemma \ref{lpro2} in a similar manner.
Since $M_{U^0}$ centralizes $N_V$ we can form the direct product $P_1:=P_{U^0}/Z=M_{U^0}T$ where
$T:=N_V/Z$ (the group $Z$ was defined in \ref{weih256}).
By construction of $U^0$
there is a lattice $V_1 \subset T$ and a
homomorphism $m:T\rightarrow M_{U^0}$
such that ${U^0/{[U^0,U^0]}}=\{m(v)v|v\in {V_1}\}\subset P_1$.
For $f\in {R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta$ (this space is defined in \ref{weih257}) we have
$$\pi^{1_\lambda,{\varphi}}(m(v)v)L(f)(yw)= \theta^m(v) {\varphi}(\tilde v) f(\tilde v^{-1}yw)\ ,$$
where $\tilde v\in N$ is any lift of $v\in T$, and $\theta^m=\theta\circ m\in {\mathcal{X}}(T)$.
\subsubsection{}
We normalize the Haar measure on $T$ such that
${\rm vol}({V_1}\backslash T)=1$.
Given a unitary character
$\vartheta\in {\mathcal{X}}(T)$,
we consider the push-down $\pi^{T,\vartheta}_*$,
which associates to $f\in C^\infty(B_Z,V_{B_Z}(1_\lambda,{\varphi}))$
the average
$$\pi^{T,\vartheta}_*(f):=\int_{T} \vartheta(z)^{-1} \pi^{1_\lambda,{\varphi}}(z) f dz$$
provided the integral converges.
Note that we define $L:{R}_{[U^0,U^0]}(1_\lambda,{\varphi})_\theta^n\rightarrow
B_{[U^0,U^0]}(1_\lambda,{\varphi})$ using a $Z$-invariant
cut-off function (recall that $Z=P_{[U^0,U^0]}$). It follows from the proof of Lemma \ref{lpro1}
that all $f\in {R}_{[U^0,U^0]}(1_\lambda,{\varphi})_\theta^n$ are $Z$-invariant.
Therefore $L(f)\in C^\infty(B_Z,V_{B_Z}(\theta,{\varphi}))$ and
$\pi^{T,\vartheta}_*(L(f))$ is well-defined (up to convergence).
\subsubsection{}
We fix an isomorphism of abelian groups $T\cong {\mathbb{R}}^{\dim(T)}$
preserving the Haar measure
such that we obtain an euclidean structure on $T$ and its Lie algebra ${\mathfrak{t}}$.
As before let ${\mathcal{X}}(T,{V_1})$ denote the set of all unitary characters of
$T$ which are trivial on ${V_1}$. We identify ${\mathcal{X}}(T)$
with $\imath {\mathfrak{t}}^*$ and ${\mathcal{X}}(T,{V_1})$ with a lattice
in $\imath {\mathfrak{t}}^*$. If $f\in {R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta$, then
by the Poisson summation formula
$$\pi^{U^0/{[U^0,U^0]}}_*(L(f))=\sum_{\vartheta\in{\mathcal{X}}(T,{V_1})}
\pi^{T,\vartheta-\theta^m}_*(L(f))$$ in the domain of convergence.
\subsubsection{}
Note that if $\theta^m\in {\mathcal{X}}(T,{V_1})$, then $\theta=0$
by the construction of $M_{U^0}$.
If $\theta\not=0$, then will show that
$\sum_{\vartheta\in{\mathcal{X}}(T,{V_1})}
\pi^{T,\vartheta-\theta^m}_*(L(f))$
gives rise to a rapidly decreasing section for all $\lambda\in {\aaaa_\C^\ast}$.
If $\theta=0$, then again $\sum_{0\not=\vartheta\in{\mathcal{X}}(T,{V_1})}
\pi^{T,\vartheta-\theta^m}_*(L(f))$ is rapidly decreasing.
The remaining term $\pi^{T,0}_*(L(f))$ contributes to
${S}_{U^0}(1_\lambda,{\varphi})\oplus L({R}_{U^0}(1_\lambda,{\varphi})^n_\theta)$
and depends meromorphically on $\lambda$.
\subsubsection{}
Observe that $A$ normalizes $Z$. This is a consequence of the assumption that the cusp is regular, and that the Langlands decomposition of $P$ is adapted (Definition \ref{t799}). It is in fact the reason for making this assumption. We see that $A$ acts on $T$.
In particular ${\mathfrak{t}}^*$ decomposes into two
eigenspaces ${\mathfrak{t}}={\mathfrak{t}}_1^*\oplus{\mathfrak{t}}_2^*$
(we write $\vartheta=\vartheta_1\oplus \vartheta_2$ for the corresponding decomposition of $\vartheta\in{\mathfrak{t}}^*$)
with respect to $A$ such that $A$ acts on ${\mathfrak{t}}_i^*$ by $a^{i\alpha}$, $i=1,2$.
We define the $A$-homogeneous "norm" on ${\mathfrak{t}}^*$ by
$|\vartheta|:=(|\vartheta_1|^4+|\vartheta_2|^2)^{1/2}$.
Furthermore we use
the euclidean structure on ${\mathfrak{t}}_i$
in order to define Laplace operators
$\Delta_i\in {\mathcal{U}}({\mathfrak{t}}_i)$ and the $A$-homogeneous operator
$\Delta:=\Delta_1^2+\Delta_2$. We fix the normalizations such that
for $\vartheta\in{\mathcal{X}}(T)$ we have
$\vartheta(\Delta^l x)=|\vartheta|^{2l}
\vartheta(x)$.
Let $\rho_{U^0/{[U^0,U^0]}}\in{\mathfrak{a}}^*$ be such that $A$ acts on
$\Lambda^{max}{\mathfrak{t}}$ by the character $a^{2\rho_{U^0/{[U^0,U^0]}}}$.
Note that $\rho_{U^0/{[U^0,U^0]}}+\rho_{[U^0,U^0]}=\rho_{U}$.
\subsubsection{}
Let $COP:{\mathcal{U}}({\mathfrak{t}})\rightarrow {\mathcal{U}}({\mathfrak{t}})\otimes {\mathcal{U}}({\mathfrak{t}})$ denote
the coproduct on ${\mathcal{U}}({\mathfrak{t}})$. There are $A_{l,j},B_{l,j},C_{l,j}\in{\mathcal{U}}({\mathfrak{t}})$
such that
$$(COP\otimes 1)\circ COP(\Delta^l)=\sum_{j} A_{l,j}\otimes B_{l,j}\otimes C_{l,j}\ .$$
Let $D\mapsto \tilde D$ be the canonical antiautomorphism of ${\mathcal{U}}({\mathfrak{t}})$.
Recall that $L(f)(xw)=\chi(xw) f(xw)$, where $\chi$ is $Z$-invariant.
By $S(T)$ we denote the unit-sphere in $T$.
For $\xi\in S(T)$ and $a\in A$ we compute
\begin{eqnarray}
&&\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f)(.w)\nonumber\\
&=&\pi^{1_\lambda,{\varphi}}(a)\pi^{1_\lambda,{\varphi}}((\Delta^l)^{a^{-1}}\xi) \pi^{1_\lambda,{\varphi}}(a^{-1})\chi f(.w)\nonumber\\
&=&a^{\rho-\lambda-4l\alpha}\pi^{1_\lambda,{\varphi}}(a)\pi^{1_\lambda,{\varphi}}(\Delta^l\xi) {\varphi}(a^{-1}) \chi(a.a^{-1}w) f(a. a^{-1}w) \nonumber \\
&=&a^{\lambda+\rho-2\rho^{[U^0,U^0]}-(n+4l)\alpha}\pi^{1_\lambda,{\varphi}}(a)\pi^{1_\lambda,{\varphi}}(\Delta^l) {\varphi}(\xi) \chi(a\xi^{-1}.a^{-1}w) f(\xi^{-1}.w)\nonumber\\
&=&a^{\lambda+\rho-2\rho^{[U^0,U^0]}-(n+4l)\alpha}\sum_{j}
\pi^{1_\lambda,{\varphi}}(a) {\varphi}(A_{l,j}\xi) \chi(a\xi^{-1}\tilde B_{l,j}.a^{-1}w) f(\xi^{-1}\tilde C_{l,j}.w)\nonumber\\
&=&
a^{2(\lambda-\rho^{[U^0,U^0]})-(n+4l)\alpha}\sum_{j}
{\varphi}(a) {\varphi}(A_{l,j}\xi) \chi(a\xi^{-1}\tilde B_{l,j}a^{-1}.w) f(\xi^{-1}\tilde C_{l,j}a^{-1}.aw)\label{zzz1}\ .
\end{eqnarray}
If $W\subset N$ is any compact subset and $D\in {\mathcal{U}}({\mathfrak{n}})$, then there
is a constant $C\in{\mathbb{R}}$ such that
$$|\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f)(yDw)|\le C a^{2(\lambda-\rho^{[U^0,U^0]})-(n+4l)\alpha+l_{\varphi}}$$
for all $\xi\in S(T)$ and $a\in A_+$.
\subsubsection{}
If ${\rm Re }(\lambda)$ is sufficiently small, then by partial integration for
$\vartheta\in {\mathcal{X}}(T,{V_1})$, $\vartheta-\theta^m\not=0$
\begin{eqnarray*}
\pi^{Z,\vartheta-\theta^m}_*(L(f))&=&\int_{Z}
(\vartheta-\theta^m)^{-1}(x)\pi^{1_\lambda,{\varphi}}(x) L(f) dx\\
&=&\frac{1}{|\vartheta-\theta^m|^{2l}}\int_{T} (\vartheta-\theta^m)^{-1}(\Delta^lx) \pi^{1_\lambda,{\varphi}}(x) L(f) dx\\
&=&\frac{1}{|\vartheta-\theta^m|^{2l}}\int_{T} (\vartheta-\theta^m)^{-1}(x)\pi^{1_\lambda,{\varphi}}(\Delta^l x) L(f) dx\\
&=&\frac{1}{|\vartheta-\theta^m|^{2l}}\int_{S(T)} \int_A
(\vartheta-\theta^m)^{-1}(\xi^a)\pi^{1_\lambda,{\varphi}}(\Delta^l \xi^a) L(f)a^{2\rho_{{U^0/{[U^0,U^0]}}}} da d\xi\ .
\end{eqnarray*}
By the estimate above
the integral converges locally uniformly on
$\{{\rm Re }(\lambda)<\rho^{U}+(n+4l)\alpha/2 -l_{\varphi}/2\}$,
and for any compact subset of this region,
$D\in{\mathcal{U}}({\mathfrak{n}})$, and compact subset $W\subset N$
there is a constant $C_1$ such that
$$|\pi^{T,\vartheta-\theta^m}_*(L(f))(yDw)|\le \frac{C_1}{ |\vartheta-\theta^m|^{2l}}$$
for all $y\in W$.
\subsubsection{}
Choosing $2l>\dim(T)$ we see that the sum
$$\sum_{\vartheta\in{\mathcal{X}}(T,{V_1}), \vartheta-\theta^m\not=0}
\pi^{T,\vartheta-\theta^m}_*(L(f))$$ converges in $C^\infty(\Omega,V(1_\lambda,{\varphi}))$.
\subsubsection{}
Refining the estimates above we now show that this
sum in fact converges in the space of rapidly decreasing functions.
Let $W_1\subset N\setminus N_V$ be a compact subset and $D\in{\mathcal{U}}({\mathfrak{n}})$.
It immediately follows from (\ref{zzz1}), that there is a constant $C_2$
such that for all $\xi\in S(T)$, $a,b\in A_+$, with $a \ge b$ and $y\in W_1$
$$|{\varphi}(b)^{-1}(\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f))(y^bDw)|\le
C_2 a^{2(\lambda-\rho^{[U^0,U^0]})-(n+4l)\alpha}| {\varphi}(ab^{-1})| \ .$$
Let $COP(D)=\sum_{h} D_h^l\otimes D_h^r$
denote the coproduct of $D$, where here
$COP:{\mathcal{U}}({\mathfrak{n}})\rightarrow {\mathcal{U}}({\mathfrak{n}})\otimes{\mathcal{U}}({\mathfrak{n}})$.
We find a constant $C_3$ such that
for all $\xi\in S(T)$, $a,b\in A_+$, with $a \le b$ and $y\in W_1$
\begin{eqnarray*}
&&|{\varphi}(b)^{-1}(\pi^{1_\lambda,{\varphi}}(\Delta^l\xi^a) L(f))(y^bDw)|\\
&=&
a^{2(\lambda-\rho^{[U^0,U^0]})-(n+4l)\alpha}|\sum_{j,h}
{\varphi}(ab^{-1}) {\varphi}(A_{l,j}\xi) \chi(a\xi^{-1}\tilde B_{l,j}a^{-1}y^bD^l_hw) f(\xi^{-1}\tilde C_{l,j}a^{-1}y^bD^r_haw)\\
&=&
a^{-4l\alpha}
b^{2(\lambda-\rho^{[U^0,U^0]})-n\alpha}
|\sum_{j,h} {\varphi}((A_{l,j}\xi)^{ab^{-1}}) \chi(a\xi^{-1}\tilde B_{l,j}a^{-1}y^bD^l_hw) f((\xi^{-1}\tilde C_{l,j})^{b^{-1}a}y(D^r_h)^{b^{-1}}w)\\
&\le& C_3 b^{2(\lambda-\rho^{[U^0,U^0]})-(n+4l)\alpha}\ .
\end{eqnarray*}
Using this estimate we see that
$${\varphi}(b)^{-1}\sum_{\vartheta\in{\mathcal{X}}(T,{V_1}),\vartheta-\theta^m\not=0}
\pi^{T,\vartheta-\theta^m}_*(L(f))(y^bDw)$$
can be estimated by $C b^{2(\lambda-\rho^{U})-(n+4l)\alpha}$,
where $C$ can be choosen uniformly for $y\in W_1$ and $\lambda$ in compact subsets
of $\{{\rm Re }(\lambda)<\rho^{U}+(n+4l)\alpha/2 -l_{\varphi}/2\}$.
Since we can choose $l$ arbitrary large we obtain
a holomorphic continuation of the sum above to all
of ${\aaaa_\C^\ast}$ and that it is rapidly decreasing with respect to $b$.
\subsubsection{}
Let now $\theta=0$ and consider
$\pi^{T,0}_*(L(f))$.
We write for ${\rm Re }(\lambda)$ sufficiently small
$$\pi^{T,0}_*(L(f))=\int_A \int_{S(T)} \pi^{1_\lambda,{\varphi}}(\xi^a) L(f) d\xi a^{2\rho_{U^0/{[U^0,U^0]}}} da\ .$$
We again employ
\begin{eqnarray*}
&&\pi^{1_\lambda,{\varphi}}(\xi^a) L(f)(yw)\\
&=&
a^{2(\lambda-\rho^{[U^0,U^0]})-n\alpha}
{\varphi}(a) {\varphi}(\xi) \chi(a\xi^{-1}y^{a^{-1}}a^{-1}w) f(\xi^{-1}a^{-1}yaw)\ .
\end{eqnarray*}
Note that
$$F(a,y):= \int_{S(T)}{\varphi}(\xi) \chi(a\xi^{-1}ya^{-1}w) f(\xi^{-1}yw) d\xi$$
is a smooth function of $y$ near $y=0$ which is independent of $a$ for large $a$.
Let
$$F(a,y^{a^{-1}})=\sum_{q=0}^r F_q(y)a^{-q\alpha} + a^{-(r+1)\alpha} R_r(a,y)$$
be the asymptotic expansion for large $a\in A$ obtained from the Taylor
series of $F(a,y)$ at $y=0$.
The remainder $R_r(a,y)$ remains bounded
as $a\to\infty$. We write
$$\pi^{T,0}_*(L(f))(yw)=I_+(y)+I_-(y)\ ,$$ where
\begin{eqnarray*}
I_+(y)&=&\int_{A_+} a^{2(\lambda-\rho^{U})-n\alpha} {\varphi}(a) F(a,y^{a^{-1}}) da\\
I_-(y)&=&\int_{A_-}\int_{S(T)} \pi^{1_\lambda,{\varphi}}(\xi^a) L(f)(yw) d\xi a^{2\rho_{U^0/{[U^0,U^0]}}} da\ .
\end{eqnarray*}
The integral $I_-$ converges for all $\lambda\in{\aaaa_\C^\ast}$
and defines a holomorphic family of smooth functions.
We write $I_+(y):=J_r^1(y)+J_r^2(y)$, where
\begin{eqnarray*}
J_r^1(y)&:=& \int_{A_+} a^{2(\lambda-\rho^U)-n\alpha} {\varphi}(a) \sum_{q=0}^r F_q(y)a^{-q\alpha} da\\
J_r^2(y)&:=& \int_{A_+} a^{2(\lambda-\rho^U)-n\alpha} {\varphi}(a)
a^{-(r+1)\alpha} R_r(a,y) da\ .
\end{eqnarray*}
The integral $J_r^2$
converges for ${\rm Re }(\lambda)<\rho^U+(n+r+1)\alpha/2-l_{\varphi}/2$
and defines a smooth function in $y$. The integral $J_r^1$ can be
evaluated:
$$J_r^1(y)=\sum_{q=0}^r (B+2(\lambda-\rho^U)-(n+q)\alpha)^{-1} F_q(y)\ .$$
It obviously defines a meromorphic family of smooth functions.
Since we can choose $r$ arbitrary large we obtain a meromorphic continuation of
$\pi^{U,1}_*(L(f))$ to all of ${\aaaa_\C^\ast}$.
\subsubsection{}
Let now $f\in {R}_{\{1\}}(1_\lambda,{\varphi})_0^n$.
If $y\in N\setminus T$ and
$b\in A$
is sufficiently large, then we have
\begin{eqnarray*}
\pi^{T,0}_*\circ L\circ [\pi^{U^0/{[U^0,U^0]}}_*](f)(y^b)&=&
\pi^{T,0}_*\circ \pi^{Z,0}_*(f)(y^b)\\
&=&\pi^{P_{U^0}}_*(f)(y^b)\\
&=&[\pi^{U^0}_*](f)(y^b)\ .
\end{eqnarray*}
We now have shown that if $f\in {R}_{[U^0,U^0]}(1_\lambda,{\varphi})^n_\theta$, then
$\pi^{U^0/[U^0,U^0]}_*(f)\in {S}_{U^0}(1_\lambda,{\varphi})\oplus L({R}_{U^0}(1_\lambda,{\varphi})^n)$,
and that $\pi^{U^0/[U^0,U^0]}_*$ depends meromorphically on $\lambda$. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
The finite group $U/U^0$ acts on ${S}_{U^0,k}(1_\lambda,{\varphi})$
and ${R}_{U^0}(1_\lambda,{\varphi})^n$. We can define
$\{\pi^{U/U^0}\}:{S}_{U^0,k}(1_\lambda,{\varphi})\rightarrow
{S}_{U,k}(1_\lambda,{\varphi})$, $\pi^{U/U^0}:B_{U^0}(1_\lambda,{\varphi})\rightarrow
B_U(1_\lambda,{\varphi})$, and $[\pi^{U/U^0}_*]:{R}_{U^0}(1_\lambda,{\varphi})^n\rightarrow
{R}_U(1_\lambda,{\varphi})^n$.
Using $\pi^{U}_*:=\pi^{U/U^0}_*\circ \pi^{U^0}_*$
and Proposition \ref{mainpure} (and its proof) we obtain
\begin{prop}\label{muncor}
The composition
$$\pi^{U}_*\circ L:{R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta\rightarrow
B_{U}(1_\lambda,{\varphi})$$
converges for ${\rm Re }(\lambda)<\rho^U+(n\alpha-l_{\varphi})/2$
and has a meromorphic continuation to all of ${\aaaa_\C^\ast}$
such that $AS\circ \pi^{U}_*\circ L = [\pi^{U}_*]$.
If $\theta\not=0$, then it is in fact
a holomorphic family of maps
$\pi^{U}_*\circ L:{R}_{\{1\}}(1_\lambda,{\varphi})^n_\theta\rightarrow
{S}_{U}(1_\lambda,{\varphi})$. If $\theta=0$,
then $\pi^{U}_*\circ L$ has at most first order poles in the set $\rho^U+\frac{n\alpha-l_{\varphi}}{2} +\frac{1}{2}{\mathbb {N}}_0$.
\end{prop}
\subsubsection{}
Combining Proposition \ref{muncor} with Lemma \ref{schwpush}
we obtain
\begin{kor}\label{u76}
Assume that the cusp associated to $U\subset P$ does not have full rank.
Then for any $k_1< k$ the push-down
$\pi_*^U: B_{\{1\},k}(1_\lambda,{\varphi}) \rightarrow B_{U,k_1}(1_\lambda,{\varphi})$
forms a meromorphic family of continuous maps
with finite-dimensional singularities
defined on $\{{\rm Re }(\lambda)<\rho^U+(k\alpha-l_{\varphi})/2\}$.
It fits into the following commutative diagram
$$
\begin{array}{ccccccccc}
0&\rightarrow&{S}_{\{1\},k}(1_\lambda,{\varphi})&\rightarrow& B_{\{1\},k}(1_\lambda,{\varphi})&\stackrel{AS}{\rightarrow}& {R}_{\{1\},k}(1_\lambda,{\varphi})&\rightarrow &0\\
& &\downarrow \{\pi^{U}_*\} & & \downarrow \pi^{U}_* & & \downarrow [\pi^{U}_*] &&\\
0&\rightarrow&{S}_{U,k}(1_\lambda,{\varphi})&\rightarrow& B_{U,k}(1_\lambda,{\varphi})& \stackrel{AS}{\rightarrow}&{R}_{U,k}(1_\lambda,{\varphi})&\rightarrow &0\ .\end{array}
$$
Moreover we have a meromorphic family of maps
$\pi_*^U:B_{\{1\}}(1_\lambda,{\varphi}) \rightarrow B_{U}(1_\lambda,{\varphi})$
with finite-dimensional singularities and defined on all of ${\aaaa_\C^\ast}$
such that
$$
\begin{array}{ccccccccc}
0&\rightarrow&{S}_{\{1\}}(1_\lambda,{\varphi})&\rightarrow& B_{\{1\}}(1_\lambda,{\varphi})&\stackrel{AS}{\rightarrow}& {R}_{\{1\}}(1_\lambda,{\varphi})&\rightarrow &0\\
& &\downarrow \{\pi^{U}_*\} & & \downarrow \pi^{U}_* & & \downarrow [\pi^{U}_*] &&\\
0&\rightarrow&{S}_{U}(1_\lambda,{\varphi})&\rightarrow& B_{U}(1_\lambda,{\varphi})& \stackrel{AS}{\rightarrow}&{R}_{U}(1_\lambda,{\varphi})&\rightarrow &0\ .\end{array}
$$
is commutative.
\end{kor}
\subsection{Push-down for cusps of full rank and for general $\sigma$} \label{maxrank}
\subsubsection{}
In this subsection $(\sigma,V_\sigma)$ denotes a Weyl-invariant
representation of $M$ as in \ref{weih241}.
Let $C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))\subset C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$
be the space of distributions which are supported on $\infty_P$.
This space can be identified $P_U$-equivariantly with the tensor product of
a generalized Verma module by $V_{\varphi}$
$$({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}$$
such that $(X\otimes s\otimes v)\in ({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}$ maps $f\in C^\infty(\partial X,V(\tilde \sigma_{-\lambda},\tilde{\varphi}))$ to $(s\otimes v)(f(Xe))$.
We can further identify
$C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))$ with $R_{\{1\}}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*$.
\subsubsection{}\label{weih121}
We now assume that the cusp associated to $U\subset P$
has full rank. In this case $N$ is the Zariski closure of $U^0$.
The space $({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}$ carries an algebraic representation
of $N$. Therefore
\begin{eqnarray*}
{}^U[({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}]&\subset&
{}^{U^0}[({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}]\\
&=&{}^N[({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}]
\end{eqnarray*}
is finite-dimensional since the space of highest weight vectors of
the ${\mathfrak{g}}$-module
${\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}}$ is finite-dimensional \cite{MR552943}.
\subsubsection{}\label{weih300}
Let ${\mathcal{E}}_{\infty_P}(\sigma,{\varphi})$ denote the sheaf of holomorphic families
$f_\nu\in{}^UC^{-\infty}(\infty_P,V(\sigma_\nu,{\varphi}))$.
Since ${\mathcal{E}}_{\infty_P}(\sigma,{\varphi})$ is torsion-free it is the space of sections
of a unique holomorphic vector bundle $E_{\infty_P}(\sigma,{\varphi})$ over ${\aaaa_\C^\ast}$.
By $E_{\infty_P}(\sigma_\lambda,{\varphi})$ we denote the fibre of $E_{\infty_P}(\sigma,{\varphi})$ at $\lambda\in{\aaaa_\C^\ast}$.
We will discuss this bundle in detail in Lemma \ref{vermaolbrich}.
\subsubsection{}
We now define the function space $B_U(\sigma_\lambda,{\varphi})$
in the case of a cusp of full rank.
\begin{ddd}
We define $R_U(\sigma_\lambda,{\varphi}):= E_{\infty_P}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*$.
Furthermore we set $$B_U(\sigma_\lambda,{\varphi}):={S}_U(\sigma_\lambda,{\varphi})\oplus {R}_U(\sigma_\lambda,{\varphi})\ .$$
Let $AS:B_U(\sigma_\lambda,{\varphi}) \rightarrow {R}_U(\sigma_\lambda,{\varphi})$
be the projection and $L:{R}_U(\sigma_\lambda,{\varphi})\rightarrow B_U(\sigma_\lambda,{\varphi})$
be the inclusion.
\end{ddd}
These families of spaces form trivial holomorphic bundles of Fr\'echet and Montel spaces over ${\aaaa_\C^\ast}$.
\subsubsection{}
\begin{ddd}\label{rolf}
We define $$[ext^U]:E_{\infty_P}(\tilde\sigma_{-\lambda},\tilde{\varphi})\hookrightarrow
{R}_{\{1\}}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*$$ as the natural inclusion.
We define the push-down
$$[\pi^U_*]:{R}_{\{1\}}(\sigma_{\lambda},{\varphi})\rightarrow
R_U(\sigma_\lambda,{\varphi})$$ to be the adjoint of $[ext^U]$.
Furthermore we define
$$(\pi^U_*)_1:C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
{S}_U(\sigma_\lambda,{\varphi})$$ by $(\pi^U_*)_1(f):=\sum_{u\in U}\pi^{\sigma_\lambda,{\varphi}}(u)(f_{|\Omega})$
(convergence provided). Finally we set
$$\pi^U_*:=(\pi^U_*)_1\oplus [\pi^U_*]:C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow B_U(\sigma_\lambda,{\varphi}) \ .$$
\end{ddd}
Note that
$[\pi^U_*]$ is a holomorphic family of surjective maps.
Once and for all we fix a right-inverse $[Q]$ of $[\pi^U_*]$.
We leave it to the interested reader to show that the
statements of Lemma \ref{compat2} hold true for cusps of full rank as well.
\subsubsection{}
\begin{lem}
The push-down
$(\pi^U_*)_1:C^\infty(\partial X,V(1_\lambda,{\varphi}))\rightarrow
{S}_U(1_\lambda,{\varphi})$ converges for ${\rm Re }(\lambda)<-l_{\varphi}/2$
and has a meromorphic continuation to all of ${\aaaa_\C^\ast}$
with finite-dimensional singularities.
\end{lem}
{\it Proof.$\:\:\:\:$}
One checks that the corresponding parts of the proofs
of Lemma \ref{schwpush} and Proposition \ref{mainpure} apply to the case
of cusps of full rank as well. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Note that in Subsection \ref{weih261} we have considered the push-down in the spherical case $\sigma=1$. We now deal with the general case using the concept of twisting.
We assume that $U\subset P$ defines a cusp of smaller rank and consider a Weyl invariant $\sigma$ (see \ref{weih241}).
We show the existence of a meromorphic family of push-down maps
$$\pi^{U}_*:C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
B_U(\sigma_\lambda,{\varphi})$$ using twisting.
We employ a finite-dimensional representation $(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$ of $G$ as in \ref{weih144}.
Note that $$\pi^U_*:C^\infty(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))
\rightarrow B_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$$
is ${\mathcal{Z}}({\mathfrak{g}})$-equivariant. Therefore we can make the following definition:
\begin{ddd}\label{cv1}
If $\lambda\not\in I_{\mathfrak{a}}$, then using Lemma \ref{compat2}, 4.,
we define the push-down
$$\pi^U_*:C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
B_U(\sigma_\lambda,{\varphi})$$
by the following coummutative diagram:
$$\begin{array}{ccccccc}
0&\rightarrow&C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))&\stackrel{i_{\sigma,\mu}}{\rightarrow}&C^\infty(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))&\stackrel{Z(\lambda)}{\rightarrow}&
C^\infty(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\\
&&\downarrow \pi^U_*&&\downarrow \pi^U_*&&\downarrow \pi^U_*\\
0&\rightarrow&B_U(\sigma_\lambda,{\varphi})&\stackrel{i^U_{\sigma,\mu}}{\rightarrow}&
{\mathcal{B}}_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})&\stackrel{Z(\lambda)}{\rightarrow}&
{\mathcal{B}}_U(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})\end{array}\ .$$
\end{ddd}
\subsubsection{}
It is clear that in the domain of convergence ${\rm Re }(\lambda)\ll 0$
this definition coincides with Definition \ref{pushdowndef}.
In order to see that
$\pi^U_*$ extends to a meromorphic family note that
$i^U_{\sigma,\mu}$ admits a meromorphic family of
left inverses $j^U_{\sigma,\mu}$ (Lemma \ref{compat2}, 5.), and that we can express
the push-down for $\sigma$ through the spherical push-down
by $j^U_{\sigma,\mu}\circ \pi^U_* \circ i^U_{\sigma,\mu}$.
\subsubsection{}
We claim that $\pi^U_*$ has finite-dimensional
singularities. Note that given $\nu\in {\mathfrak{a}}^*$
there exists $k\in{\mathbb {N}}_0$ such that the restriction of
$\pi^U_*$ to $S_{\{1\},k}(\sigma_\lambda,{\varphi})\cap B_{\{1\}}(\sigma_\lambda,{\varphi})$
converges for all ${\rm Re }(\lambda)<\nu$.
The rank of the singularities of
$\pi^U_*$ for ${\rm Re }(\lambda)<\nu$ is bounded
by the codimension of
$S_{\{1\},k}(\sigma_\lambda,{\varphi})\cap B_{\{1\}}(\sigma_\lambda,{\varphi})$
in $B_{\{1\}}(\sigma_\lambda,{\varphi})$ which is finite. This proves the claim.
\subsubsection{}
The following corollary is a consequence of the discussion
above and Corollary \ref{u76}.
\begin{kor}\label{neuj106}
Let $U\subset P$ define a regular cusp and ${\varphi}$ be an admissible twist.
Then the push-down is a meromorphic family of
continuous maps $\pi_*^U:B_{\{1\}}(\sigma_\lambda,{\varphi}) \rightarrow B_{U}(\sigma_\lambda,{\varphi})$
with finite-dimensional singularities such that
$$
\begin{array}{ccccccccc}
0&\rightarrow&{S}_{\{1\}}(\sigma_\lambda,{\varphi})&\rightarrow& B_{\{1\}}(\sigma_\lambda,{\varphi})&\stackrel{AS}{\rightarrow}& {R}_{\{1\}}(\sigma_\lambda,{\varphi})&\rightarrow &0\\
& &\downarrow \{\pi^{U}_*\} & & \downarrow \pi^{U}_* & & \downarrow [\pi^{U}_*] &&\\
0&\rightarrow&{S}_{U}(\sigma_\lambda,{\varphi})&\rightarrow& B_{U}(\sigma_\lambda,{\varphi})& \stackrel{AS}{\rightarrow}&{R}_{U}(\sigma_\lambda,{\varphi})&\rightarrow &0\ .\end{array}
$$
is commutative.
\end{kor}
\subsection{Compatibility with embedding}\label{weih116}
\subsubsection{}
In this subsection we assume that $G^n$ belongs to the list
$$\{Spin(1,n), SO(1,n)_0, SU(1,n), Sp(1,n)\}\ .$$
First assume that $U\subset P$ defines a cusp of smaller rank.
Then we have a commutative diagram
$$\begin{array}{ccc} A_{\{1\}}(1^{n+1}_{\lambda},{\varphi})^m&\stackrel{i^*}{\rightarrow}
&A_{\{1\}}(1^{n}_{\lambda-\zeta},{\varphi})^m\\
\downarrow \pi^{P^{n+1}_U}_*&&\downarrow \pi^{P^n_U}_*\\
A_{P_U}(1^{n+1}_{\lambda},{\varphi})^m&\stackrel{i_U^*}{\rightarrow}
&A_{P_U}(1^{n}_{\lambda-\zeta},{\varphi})^m
\end{array}\ .$$
In fact, commutativity is obvious in the domain of convergence.
\subsubsection{}
By the definition of the spaces
$R_{\{1\}}(1^{n+1}_{\lambda},{\varphi})$, $R_{\{1\}}(1^{n}_{\lambda-\zeta},{\varphi})$
as the spaces of asymptotics of smooth sections and the fact that
that $i^*$ maps smooth sections to smooth sections we obtain $U$-equivariant maps $R_{\{1\}}(1^{n+1}_{\lambda},{\varphi})^m\stackrel{i^*}{\rightarrow}
R_{\{1\}}(1^{n}_{\lambda-\zeta},{\varphi})^m$, $m\in{\mathbb {N}}_0$.
It follows from the naturality of the image bundle Lemma \ref{bunle}
that we have a commutative diagram
$$\begin{array}{ccc} R_{\{1\}}(1^{n+1}_{\lambda},{\varphi})^m&\stackrel{i^*}{\rightarrow}
&R_{\{1\}}(1^{n}_{\lambda-\zeta},{\varphi})^m\\
\downarrow [\pi^{U,n+1}_*]&&\downarrow[\pi^{U,n}_*]\\
R_U(1^{n+1}_{\lambda},{\varphi})^m&\stackrel{i_U^*}{\rightarrow}
&R_{U}(1^{n}_{\lambda-\zeta},{\varphi})^m
\end{array}\ .$$
\subsubsection{}
On the level of Schwartz spaces we have for all $k\in{\mathbb {N}}_0$
$$\begin{array}{ccc} S_{\{1\},k}(1^{n+1}_{\lambda},{\varphi})&\stackrel{i^*}{\rightarrow}
&S_{\{1\},k}(1^{n}_{\lambda-\zeta},{\varphi})\\
\downarrow \{\pi^{U,n+1}_*\}&&\downarrow\{\pi^{U,n}_*\}\\
S_{U,k}(1^{n+1}_{\lambda},{\varphi})&\stackrel{i_U^*}{\rightarrow}
&S_{U,k}(1^{n}_{\lambda-\zeta},{\varphi})
\end{array}\ .$$
In order to see that $i^*$ and $i^*_U$ (initially defined on spaces of smooth sections)
induce maps between Schwartz spaces one checks that these maps are bounded with respect to the norms
$\|.\|_{k,d}$ introduced in \ref{weih259}. Commutativity of the diagram is clear.
\subsubsection{}
We now easily obtain the diagram
$$\begin{array}{ccc} B_{\{1\},k}(1^{n+1}_{\lambda},{\varphi})&\stackrel{i^*}{\rightarrow}
&B_{\{1\},k}(1^{n}_{\lambda-\zeta},{\varphi})\\
\downarrow \pi^{U,n+1}_*&&\downarrow\pi^{U,n}_* \\
B_{U,k}(1^{n+1}_{\lambda},{\varphi})&\stackrel{i_U^*}{\rightarrow}
&B_{U,k}(1^{n}_{\lambda-\zeta},{\varphi})
\end{array}\ .$$
It gives Proposition \ref{rrttee} in the case of a pure cusp of lower rank.
\subsubsection{}
We now assume that $U\subset P^n$ defines a cusp of full rank.
We are going to construct a map
$i_U^*:B_U(1^{n+1}_{\lambda},{\varphi})\rightarrow
B_U(1^n_{\lambda-\zeta},{\varphi})$ as the direct sum of
$$(i_U^*)_1:B_U(1^{n+1}_{\lambda },{\varphi})\rightarrow
S_U(1^{n}_{\lambda },{\varphi})$$ and
$$(i_U^*)_2:B_U(1^{n+1}_{\lambda },{\varphi})\rightarrow
R_U(1^{n}_{\lambda-\zeta},{\varphi})\ .$$
While $(i_U^*)_1$ is just the restriction of sections
the definition of $(i_U^*)_2$ is more complicated.
In order to verify its properties we need some results of Subsection \ref{samel}.
\subsubsection{}
In the following we discuss $(i_U^*)_2$.
This map comes as a meromorphic family and will essentially be fixed
by the condition
\begin{equation}\label{tobsucht}\pi^{U,n}_*\circ i^*= i_U^*\circ \pi^{U,n+1}_*\ .\end{equation}
The details are as follows.
\subsubsection{}
Since $E_{\infty_P}(1^{n}_{-\lambda+\zeta},\tilde{\varphi})$ (see \ref{weih300} for notation) is finite-dimensional it consists of distributions of uniformly bounded order. Further, since these distributions are supported in $\infty_P$
we can choose
$k\in {\mathbb {N}}$ be such that
$E_{\infty_P}(1^{n}_{-\mu+\zeta},\tilde{\varphi})$ pairs trivially with the space ${S}_{\{1\},k-2\rho_U}(1^n_{\mu-\zeta}, {\varphi})$ for all $\mu$ in some compact neighbourhood of $\lambda$. Thus we have an inclusion
$E_{\infty_P}(1^{n}_{-\mu+\zeta},\tilde{\varphi})\subset{R}_{\{1\},k}(1^n_{\mu-\zeta}, {\varphi})^*$
(see Definition \ref{weih231} for notation and use the dual of (\ref{weih210})).
Note that the target of $(i_U^*)_2$ is the dual of $E_{\infty_P}(1^{n}_{-\lambda+\zeta},\tilde{\varphi})$.
We define $$(i^*_U)_2: B_U(1^{n+1}_{\lambda }, {\varphi}) \rightarrow
{R}_U(1^n_{ \lambda-\zeta},{\varphi})$$
by the condition that
\begin{equation}\label{weih404}
\langle\phi,(i^*_U)_2(f)\rangle=\langle i_*(\phi), L\circ [Q]\circ AS(f)\rangle
\end{equation}
for all $\phi\in E_{\infty_P}(1^{n}_{-\lambda+\zeta},\tilde{\varphi})$, where
$i_*:C^{-\infty}(\partial X^n,V(1^n_{-\lambda+\zeta},\tilde{\varphi}))\rightarrow C^{-\infty}(\partial X^{n+1},V(1^{n+1}_{-\lambda },\tilde{\varphi}))$
is the natural inclusion adjoint to $i^*$ (see \ref{weih301}),
$L$ is the split of (\ref{weih210}), $[Q]$ is defined in \ref{weih500}, and
$AS(f)\in {R}_{\{1\},k}(1^n_{\lambda}, {\varphi})$.
This formula defines $(i^*_U)_2(f)$ for generic $\lambda\in {\aaaa_\C^\ast}$ where $[Q]$ is regular.
The meromorphic family is given by
$$(i^*_U)_2:=(i_*)_{|E_{\infty_P}(1^{n}_{-\lambda+\zeta},\tilde{\varphi})}^*\circ L\circ [Q]\circ AS\ .$$
\subsubsection{}
\begin{lem}\label{sucht}
The definition of $(i^*_U)_2$ is independent of the choice of $k$, the split $L$ and $[Q]$.
\end{lem}
{\it Proof.$\:\:\:\:$}
Let $L^\prime\circ [Q^\prime]\circ AS^\prime$ be defined with different choices (assume that $k^\prime\ge k$).
Let $f_\mu\in B_U(1^{n+1}_\mu,{\varphi})$ be the germ of a holomorphic family near $\lambda$, and consider a family $\phi_\mu\in E_{\infty_P}(1^{n}_{-\mu+\zeta},\tilde{\varphi})$.
Then we have
\begin{eqnarray*}
\lefteqn{AS\circ \pi_*^{U,n+1}(L^\prime\circ [Q^\prime]\circ AS^\prime(f)- L\circ [Q]\circ AS(f))}&&\\&\stackrel{Cor. \ref{u76}}{=}&
[\pi_*^{U,n+1}]\circ AS (L^\prime\circ [Q^\prime]\circ AS^\prime(f)-L\circ [Q]\circ AS(f))\\
&=&[\pi_*^{U,n+1}] ([Q^\prime]\circ AS^\prime(f)-[Q]\circ L\circ AS(f))\\
&=&AS^\prime(f)-AS(f)\ .
\end{eqnarray*}
We define (see Lemma \ref{schwspl} for $\{Q\}$)
\begin{eqnarray*}
\Delta&:=&L^\prime\circ [Q^\prime]\circ AS^\prime(f)+\{Q\}\circ L(AS^\prime(f)-AS(f))\\&-& L\circ [Q]\circ AS(f)\ .
\end{eqnarray*}
Then by construction we have
\begin{equation}\label{weih402}
\pi_*^{U,n+1}(\Delta)=0\ .
\end{equation}
Using results which we will prove (independently of the present stuff) in Subsection \ref{samel}
we show that $\langle i_*(\phi),\Delta\rangle=0$. Proposition \ref{maertins} states an equality
of two spaces defined in \ref{weih400} and \ref{weih401}:
$${\mathrm{ Ext}}_U(1^{n+1}_{-\mu},\tilde{\varphi})=E_U(1^{n+1}_{-\mu},\tilde{\varphi})\ .$$
On the one hand, $$E_U(1^{n+1}_{-\mu},\tilde{\varphi})\subset {}^UC^{-\infty}(\partial X^{n+1},V(1^{n+1}_{-\mu},\tilde {\varphi}))$$ is the space of evaluations at $\mu$ of holomorphic families of
$U$-invariant distributions. In particular we have $i_*(\phi_\mu)\in E_U(1^{n+1}_{-\mu},\tilde{\varphi})$.
On the other hand, for generic $\mu$ the space ${\mathrm{ Ext}}_U(1^{n+1}_{-\mu},\tilde{\varphi})$ is contained in the annihilator of the kernel of $\pi^{U,n+1}_*$. Using (\ref{weih402}) we obtain
$\langle i_*(\phi_\mu),\Delta_\mu\rangle=0$.
By our choice of $k$ we have
$\langle i_*(\phi_\mu),g\rangle=0$ for every $g\in {S}_{\{1\},k-2\rho_U}(1^{n+1}_{\mu}, {\varphi})$ and $\mu$ near $\lambda$
since $g$ vanishes at $\infty_P$ with order larger than $k$.
Note that
$L(AS^\prime(f_\mu)-AS(f_\mu))\in {S}_{U,k}(1^{n+1}_{\mu}, {\varphi})$.
Since $\{Q\}\circ L(AS^\prime(f_\mu)-AS(f_\mu))\in {S}_{\{1\},k-2\rho_U}(1^{n+1}_{\mu}, {\varphi})$
it follows that
$$\langle i_*(\phi_\mu), \{Q\}\circ L(AS^\prime(f_\mu)-AS(f_\mu))\rangle=0$$
for generic $\mu$ near $\lambda$.
Finally we conclude for these $¸\mu$ that
\begin{eqnarray*}
\lefteqn{
\langle i_*(\phi_\mu), L^\prime\circ [Q]^\prime\circ AS^\prime(f_\mu)\rangle}&&\\&=&
\langle i_*(\phi_\mu), L^\prime\circ [Q^\prime]\circ AS^\prime(f_\mu)+\{Q\}\circ L(AS^\prime(f_\mu)-AS(f_\mu))\rangle\\
&=&\langle i_*(\phi_\mu), L\circ [Q]\circ AS(f_\mu)\rangle\ .\end{eqnarray*}
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
\begin{lem}
The equality (\ref{tobsucht}) holds true.
\end{lem}
{\it Proof.$\:\:\:\:$}
Let $f\in B_{\{1\}}(1^{n+1}_{\lambda }, {\varphi})$.
Then we have
\begin{eqnarray}\lefteqn{
\pi_*^{U,n+1}(f-L\circ [Q]\circ AS\circ \pi^{U,n+1}_*(f)}&&\nonumber\\&&-\{Q\}(\pi^{U,n+1}_*(f)- \pi^{U,n+1}_*\circ L\circ [Q]\circ AS\circ \pi^{U,n+1}_*(f))\, =\,0\ .\label{tob}
\end{eqnarray}
We have seen in the proof of Lemma \ref{sucht} that $i_*(\phi)$ annihilates the kernel of $\pi^{U,n+1}_*$
(for generic $\lambda$, where all the maps are regular). We combine this fact with (\ref{tob}) in order to derive Equality (\ref{entscheident}) below.
For all $\phi\in E_{\infty_P}(1^{n}_{-\lambda+\zeta},\tilde{\varphi})$ we have
\begin{eqnarray}
\langle\phi, AS\circ i^*_U\circ \pi^{U,n+1}_*(f)\rangle&\stackrel{def}{=}&
\langle\phi, (i^*_U)_2 \circ \pi_*^{U,n+1}(f)\rangle\nonumber\\
&\stackrel{(\ref{weih404})}{=}&\langle i_*(\phi),L\circ [Q]\circ AS\circ \pi^{U,n+1}_*(f)\rangle \nonumber\\
&=&\langle i_*(\phi),L\circ [Q]\circ AS\circ \pi^{U,n+1}_*(f)\rangle \nonumber\\
&&+\{Q\}(\pi^{U,n+1}_*(f)-\pi^{U,n+1}_*\circ L\circ [Q]\circ AS\circ \pi^{U,n+1}_*(f))\rangle\nonumber\\
&=&\langle i_*(\phi),f \rangle \label{entscheident}\\
&=&\langle \phi,i^*(f) \rangle \nonumber\\
&=&\langle \phi,[\pi^{U,n}_*]\circ AS\circ i^*(f) \rangle\nonumber\ .
\end{eqnarray}
Since clearly $\{\pi^{U,n}_*\}\circ i^*=(i_U^*)_1\circ \{\pi^{U,n+1}_*\}$
we conclude the required indentity (\ref{tobsucht}). \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Let us combine the results of the present subsection into one statement.
Let now $U\subset P^n$ define a cusp of arbitrary rank.
Note that $(i_U^*)_2$ and therefore $i^*_U$ may have poles in the case of a
cusp of full rank. The following Proposition settles Proposition \ref{rrttee} in the case of pure cusps.
\begin{prop}\label{klopp}
We have the following commutative diagram (to be understood as an identity of meromorphic families if $i_U^*$ has poles)
\begin{equation}\label{weih502}
\begin{array}{ccc}
C^\infty(\partial X,V(1^{n+1}_{\lambda },{\varphi})) &\stackrel{i^*}{\rightarrow}&C^\infty(\partial X^n, V(1^n_{\lambda-\zeta},{\varphi}))\\
\downarrow \pi^{U,n+1}_*&&\downarrow \pi^{U,n}_*\\
B_U(1^{n+1}_{\lambda },{\varphi})&\stackrel{i_U^*}{\rightarrow }& B_U(1^n_{\lambda-\zeta},{\varphi})
\end{array}\ .
\end{equation}
\end{prop}
\subsection{Extension and restriction}\label{samel}
\subsubsection{}
In the present subsection we assume for simplicity that the twist
${\varphi}$ is normalized such that all its highest $A$-weights are zero.
This differs from the convention adopted in \ref{weih213}. Our present convention has the effect that
the dual $\tilde {\varphi}$ has the normalization adopted in \ref{weih213}.
Let $l_{\varphi}$ be the highest weight of $\tilde{{\varphi}}$,
i.e. $l_{\varphi}:=l_{\tilde{{\varphi}}}$.
\subsubsection{}
The space $C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))$
carries an action $\pi^{\sigma_\lambda,{\varphi}}$ of $P_UA$.
Using the isomorphism of $P_UA$-modules $$C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))\cong
({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}$$
and of $M_UA$-modules (compare \ref{neuj100})
$$({\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}})\otimes V_{\varphi}
\cong {\mathcal{U}}(\bar{\mathfrak{n}})\otimes V_{\sigma_{\lambda+2\rho}}\otimes V_{\varphi}$$
given by the PBW-theorem we see that $A$ acts semisimply.
Let $C^{-\infty}(\infty_P,V(\sigma,{\varphi}))^n$ denote
the subspace on which $A$ acts with weight $-\lambda-\rho-n\alpha$.
Then we have $$C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))^n=
(R_{\{1\}}(\tilde\sigma_{-\lambda},\tilde{\varphi})^n)^*\ .$$
\subsubsection{}
Since the action of $P_U$ is algebraic
and $U\subset P_U$ is Zariski dense we have
$${}^U C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))=
{}^{P_U} C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))\ .$$
Since $A$ normalizes $P_U$ we
conclude that ${}^U C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))$
is an $A$-invariant subspace.
In particular, we obtain a decomposition
$${}^U C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))=\bigoplus_{n=0}^\infty
{}^U C^{-\infty}(\infty_P,V(\sigma_\lambda,{\varphi}))^n= \bigoplus_{n=0}^\infty{}^U (R_{\{1\}}(\tilde\sigma_{-\lambda},\tilde{\varphi})^n)^*\ .$$
\subsubsection{}\label{neuj703}
Let ${\mathcal{E}}_{\infty_P}(\sigma,{\varphi})^n$ be the sheaf
of holomorphic families $f_\nu\in {}^U (R_{\{1\}}(\tilde\sigma_{-\nu},\tilde{\varphi})^n)^*$.
This sheaf is torsion-free, and it is therefore the sheaf of holomorphic
sections of a holomorphic vector bundle $E_{\infty_P}(\sigma,{\varphi})^n$.
By $E_{\infty_P}(\sigma_\lambda,{\varphi})^n$ we denote its fibre at $\lambda$.
For each $n\in{\mathbb {N}}_0$
we define the space
$\bar Q_(\sigma_\lambda,{\varphi})^n$ by the exact sequence
$$0\rightarrow E_{\infty_P}(\sigma_\lambda,{\varphi})^n\rightarrow
{}^U(R_{\{1\}}(\tilde\sigma_{-\lambda},\tilde{\varphi})^n)^*\rightarrow
\bar Q_{\infty_P}(\sigma_\lambda,{\varphi})^n\rightarrow 0\ .$$
Furthermore let $E_{\infty_P}(\sigma_\lambda,{\varphi}):=\bigoplus_{n=0}^\infty E_{\infty_P}(\sigma_\lambda,{\varphi})^n$
and $\bar Q_{\infty_P}(\sigma_\lambda,{\varphi}):=\bigoplus_{n=0}^\infty \bar Q_{\infty_P}(\sigma_\lambda,{\varphi})^n$.
\begin{ddd}\label{neuj107}
We call the elements of $E_{\infty_P}(\sigma_\lambda,{\varphi})^n$ deformable. An element of\linebreak[4]
${}^U(R_{\{1\}}(\tilde\sigma_{-\lambda},\tilde{\varphi})^n)^*$ is called undeformable, if it represents a nontrivial class in $\bar Q_{\infty_P}(\sigma_\lambda,{\varphi})^n$.
\end{ddd}
\subsubsection{}
Let ${\mathfrak{t}}\subset {\mathfrak{m}}$ be a Cartan subalgebra of ${\mathfrak{m}}$. Then
${\mathfrak{h}}:={\mathfrak{t}} \oplus {\mathfrak{a}}$ is a Cartan subalgebra of ${\mathfrak{g}}$.
By $\Delta^+({\mathfrak{g}},{\mathfrak{h}})$ we denote a positive root system
which is compatible with the orientation of ${\mathfrak{a}}$.
By $\Delta^+({\mathfrak{m}},{\mathfrak{h}})\subset \Delta^+({\mathfrak{g}},{\mathfrak{h}})$
we denote the subsystem of roots of ${\mathfrak{m}}$.
For each $\sigma\in \hat{M}$ we define
\begin{equation}\label{neuj103}
{\mathfrak{a}}^*\ni d(\sigma):=-\rho+\max\{
\frac{\langle \mu_\sigma,\varepsilon\rangle}{\langle \alpha,\varepsilon\rangle}
\:|\:\varepsilon\in \Delta^+({\mathfrak{g}},{\mathfrak{h}})\setminus \Delta^+({\mathfrak{m}},{\mathfrak{h}})\} \alpha \ ,
\end{equation}
where $\mu_\sigma$ is the highest weight of $\sigma$, and
$\langle.,.\rangle$ is any Weyl-invariant scalar product on
${\mathfrak{h}}$.
There is a natural action of $P_UA$ on the space
${\rm Pol}(N,V_{\sigma^w_{-\lambda}}\otimes V_{\varphi})$ of polynomials on $N$ with values in $V_{\sigma^w_{-\lambda}}\otimes V_{\varphi}$ (compare \ref{neuj101})
given by
$$(man.f)(x) = (\sigma^w_{-\lambda}\otimes {\varphi})(man) f((n^{-1}x)^{m^{-1}a^{-1}}),\quad
m\in M_U,a\in A, n\in N_V\ .$$
\subsubsection{}
\begin{lem}\label{vermaolbrich}
\begin{enumerate}
\item
There is a holomorphic family of $A$-equivariant
maps
$$j_\lambda : E_{\infty_P}(\sigma_\lambda,{\varphi}){\rightarrow}\ {}^{P_U}{\rm Pol}(N,V_{\sigma^w_{-\lambda}}\otimes V_{\varphi}) \ .$$
\item
If ${\rm Re }(\lambda)> d(\sigma)$ or $\lambda\not\in I_{\mathfrak{a}}$,
then $j_\lambda$ is an isomorphism and $\bar Q_{\infty_P}(\sigma_\lambda,{\varphi})=0$.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
\subsubsection{}\label{gaa100}
Let
$$\hat{J}^w_\lambda:C^{-\infty}(\partial X,V(\sigma_\lambda))
\rightarrow C^{-\infty}(\partial X,V(\sigma^w_{-\lambda}))$$
be the unnormalized Knapp-Stein intertwining operator
(compare \cite{MR1749869}, Sec. 5, (15)).
In order to fix the conventions we recall its definition.
The restriction of $\hat{J}^w_\lambda$ to smooth sections is given for
${\rm Re }(\lambda)<0$ by
$$(\hat J^w_{\lambda})f(g)=\int_{\bar N} f(gw\bar n) d\bar n\ .$$
For the rest of parameters it is defined by meromorphic continuation, and it extends by continuity
to distributions.
\subsubsection{}
By $$j_\lambda:C^{-\infty}(\infty_P,V(\sigma_\lambda))
\rightarrow {\rm Pol}(N,V_{\sigma^w_{-\lambda}})$$
we denote the off-diagonal part of the Knapp-Stein intertwining operator.
Here we identify
${\rm Pol}(N,V_{\sigma^w_{-\lambda}})$ with a subspace
of $C^\infty(\Omega_P,V(\sigma^w_{-\lambda}))$
such that $p\in {\rm Pol}(N,V_{\sigma^w_{-\lambda}})$
corresponds to $f_p\in C^\infty(\Omega_P,V(\sigma^w_{-\lambda}))$
with $f_p(xw)=p(x)$.
The off-diagonal part of $\hat{J}^w_\lambda$ maps to polynomials since it is $P$-equivariant and the elements of $C^{-\infty}(\infty_P,V(\sigma_{\lambda}))$ are $P$-finite.
Alternatively, using the identification
$$C^{-\infty}(\infty_P,V(\sigma_{\lambda}))\cong\linebreak[4] {\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}}\ ,$$
we can write
$$j_\lambda(X\otimes v)(n)=\pi^{\sigma^w_{-\lambda}}(X) f_{1_v}(nw)\ ,$$
where $1_v\in {\rm Pol}(N,V_{\sigma^w_{-\lambda}})$ is the constant
polynomial with value $v\in V_{\sigma^w_{-\lambda}}$.
\subsubsection{}
The map $j_\lambda$ is in fact ${\mathfrak{g}}$-equivariant, where
the action $\pi^{\sigma^w_{-\lambda}}$ of ${\mathfrak{g}}$
on ${\rm Pol}(N,V_{\sigma^w_{-\lambda}})$ is induced by the
embedding ${\rm Pol}(N,V_{\sigma^w_{-\lambda}})\subset C^\infty(\Omega_P,V(\sigma^w_{-\lambda}))$.
Therefore $\ker(j_\lambda)$ is a ${\mathfrak{g}}$-submodule
of the Verma module ${\mathcal{U}}({\mathfrak{g}})\otimes_{{\mathcal{U}}({\mathfrak{p}})} V_{\sigma_{\lambda+2\rho}}$.
By \cite{MR552943}, Satz 1.17, this Verma module is irreducible
for ${\rm Re }(\lambda)> d(\sigma)$. If $\lambda\not\in I_{\mathfrak{a}}$,
then $j_\lambda$ is injective by \cite{MR1749869}, Lemma 6.7.
\subsubsection{}
Since $$\dim {\rm Pol}(N,V_{\sigma^w_{-\lambda}})^n= \dim C^{-\infty}(\infty_P,V(\sigma_\lambda))^n = \dim ({\mathcal{U}}(\bar{\mathfrak{n}})\otimes V_{\sigma_\lambda})^n\ ,$$ we conclude that
$j_\lambda$ is in fact an isomorphism.
After tensoring with $V_{\varphi}$ and taking $U$-invariants
we obtain an isomorphism
$$j_\lambda:{}^U(R_{\{1\}}(\tilde\sigma_{-\lambda},\tilde{\varphi})^n)^* \cong {}^{P_U}{\rm Pol}(N,V_{\sigma^w_{-\lambda}}\otimes V_{\varphi})^n\ .$$
Since $j_\lambda$ and its inverse depend holomorphically on $\lambda$
we have $${}^{P_U}{\rm Pol}(N,V_{\sigma^w_{-\lambda}}\otimes V_{\varphi})^n\cong E_{\infty_P}(\sigma_\lambda,{\varphi})^n$$ for $\lambda\in{\aaaa_\C^\ast}$
with ${\rm Re }(\lambda) > d(\sigma)$ or $\lambda\not\in I_{\mathfrak{a}}$.
This proves the lemma. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Recall the definition of the function spaces $B_{U,k}(\tilde\sigma_{-\lambda},\tilde{\varphi})$ given in \ref{neuj104}.
\begin{ddd}\label{neuj1002}
We define
$D_{U,k}(\sigma_\lambda,{\varphi})$ to be the dual space to
$B_{U,k}(\tilde\sigma_{-\lambda},\tilde{\varphi})$. Furthermore let
$D_{U}(\sigma_\lambda,{\varphi}):=\bigcup_{k\in{\mathbb {N}}_0} D_{U,k}(\sigma_\lambda,{\varphi})$.
\end{ddd}
As a consequence of the corresponding properties of the family of function spaces
the family of spaces $D_{U,k}(\sigma_\lambda,{\varphi})$, $\lambda\in{\aaaa_\C^\ast}$, forms local trivial holomorphic bundles of dual Fr\'echet spaces. Furthermore, the spaces
$D_{U}(\sigma_\lambda,{\varphi})$ are Montel and form a direct limit
of locally trivial holomorphic bundles.
\subsubsection{}
Recall the definition of the push-down Definition \ref{pushdowndef}.
Its meromorphic continuation was finally established in Corollary \ref{neuj106}.
\begin{ddd}
We define the extension map
$$ext^U:D_{U}(\sigma_\lambda,{\varphi})\rightarrow D_{\{1\}}(\sigma_\lambda,{\varphi})$$
as the adjoint of the push-down $$\pi^U_*:B_{\{1\}}(\tilde\sigma_{-\lambda},\tilde {\varphi})
\rightarrow B_{U}(\tilde\sigma_{-\lambda},\tilde {\varphi})\ .$$
\end{ddd}
It follows from the corresponding properties of the push-down that
the extension maps form a meromorphic family of
continuous maps with finite-dimensional singularities.
\subsubsection{}
In the remainder of the present subsection we discuss the case
$\sigma=1$.
Assume that the cusp associated to $U\subset P$ has lower rank.
For $k\in {\mathbb {N}}_0$ and ${\rm Re }(\lambda)>-\rho^U+(l_{\varphi}-k\alpha)/2$ such that $$\pi^U_*:B_{\{1\},k}(1_{-\lambda},\tilde {\varphi})
\rightarrow B_{U,k}(1_{-\lambda},\tilde {\varphi})$$
is regular (compare Corollary \ref{u76}) we also have a map
$$ext^U:D_{U,k}(1_\lambda,{\varphi})\rightarrow D_{\{1\},k}(1_\lambda,{\varphi})$$
defined as the adjoint $\pi^U_*$. If $k_1 > k$,
then
$ext^U:D_{U,k}(1_\lambda,{\varphi})\rightarrow D_{\{1\},k_1}(1_\lambda,{\varphi})$
is a meromorphic family
of continuous maps with finite-dimensional singularities
defined for
${\rm Re }(\lambda)>-\rho^U+(l_{\varphi}-k\alpha)/2$.
\subsubsection{}\label{weih400}
\begin{ddd}\label{neuj108}
We define $$Ext_{U}(1_\lambda,{\varphi})\subset D_{\{1\}}(1_\lambda,{\varphi})$$ to
be the subspace of all $f\in D_{\{1\}}(1_\lambda,{\varphi})$ of the form
$ext^U(h)_\lambda$, where $h_\mu\in D_U(1_\mu,{\varphi})$ is
a meromorphic family defined near $\lambda$ such that $\mu\mapsto ext^U(h)_\mu$ is
regular at $\mu=\lambda$. In a similar manner we define
$Ext_{U,k}(1_\lambda,{\varphi})$ to by requiring in addition that $h_\mu\in D_{U,k}(1_\mu,{\varphi})$.
\end{ddd}
The subspaces $Ext_{U,k}(1_\lambda,{\varphi})\subset Ext_{U}(1_\lambda,{\varphi})$
are defined for ${\rm Re }(\lambda)>-\rho^U+(l_{\varphi}-k\alpha)/2$.
The space $Ext_{U}(1_\lambda,{\varphi})$ plays the role of the range of the extension.
\subsubsection{}\label{neuj471}
It is clear that $$Ext_{U}(1_\lambda,{\varphi})\subset {}^U D_{\{1\}}(1_\lambda,{\varphi})\ .$$
In order to describe to which extent the space ${}^U D_{\{1\}}(1_\lambda,{\varphi})$ is exhausted by $Ext_{U}(1_\lambda,{\varphi})$
we define $Q_{U}(1_\lambda,{\varphi})$ to be the following quotient:
$$0\rightarrow Ext_{U}(1_\lambda,{\varphi}) \rightarrow {}^U D_{\{1\}}(1_\lambda,{\varphi})
\rightarrow Q_{U}(1_\lambda,{\varphi})\rightarrow 0\ .$$
The elements in the space $Q_{U}(1_\lambda,{\varphi})$ turn out to be somewhat uncontrollable.
We therefore take much effort to show that this space is trivial under certain
conditions.
\subsubsection{}\label{weih401}
We now define the space of deformable $U$-invariant distributions (compare
Def.~\ref{neuj107} for a similar definition with an additional support condition).
\begin{ddd}\label{neuj1088}
We define the subspace $$E_U(1_\lambda,{\varphi})\subset {}^UD_{\{1\}}(1_\lambda,{\varphi})$$
as the space of evaluations of germs at $\lambda$ of holomorphic families $f_\nu\in {}^UD_{\{1\}}(1_\nu,{\varphi})$. In a similar manner we define
$E_{U,k}(1_\lambda,{\varphi})\subset {}^UD_{\{1\},k}(1_\lambda,{\varphi})$ as the subspace of evaluations of families with the additional property that $f_\nu\in{}^UD_{\{1\},k}(1_\nu,{\varphi})$.
\end{ddd}
\subsubsection{}
We define the space $\bar Q_{U}(1_\lambda,{\varphi})$ as the quotient
$$0\rightarrow E_{U}(1_\lambda,{\varphi}) \rightarrow {}^U D_{\{1\}}(1_\lambda,{\varphi})
\rightarrow \bar Q_{U}(1_\lambda,{\varphi})\rightarrow 0\ .$$
\begin{ddd}
An element of ${}^U D_{\{1\}}(1_\lambda,{\varphi})$
which represents a non-trivial class in $\bar Q_{U}(1_\lambda,{\varphi})$
is called undeformable.
\end{ddd}
(compare Definition \ref{neuj107}).
\subsubsection{}
It follows immediately from the definitions that
$$Ext_{U}(1_\lambda,{\varphi}) \subset E_U(1_\lambda,{\varphi})\ .$$
Hence we have a surjection
$$Q_U(1_\lambda,{\varphi})\to \bar Q_U(1_\lambda,{\varphi})\ .$$
One of the goals of the present subsection is to show
that
$$Ext_U(1_\lambda,{\varphi})=E_U(1_\lambda,{\varphi})\ .$$
Furthermore, we want to show that for many (generic) $\lambda\in {\aaaa_\C^\ast}$
every element of ${}^U D_{\{1\}}(1_\lambda,{\varphi})$
is deformable, i.e. $Q_{U}(1_\lambda,{\varphi})\cong 0$.
The results are stated in Proposition \ref{maertins}.
\subsubsection{}
Now we come to the definition of a left-inverse of $ext^U$: the restriction map $res^U$.
We fix $k\in {\mathbb {N}}_0$.
Recall the construction of the meromorphic family of right-inverses of $[\pi_*^U]$
$$[Q]:R_{U,k}(1_\lambda,\tilde{{\varphi}})\to R_{\{1\},k}(1_\lambda,\tilde {\varphi})$$
from \ref{weih500}. Let $k_1\in{\mathbb {N}}_0$ be such that $k_1 \alpha<k\alpha-2\rho_U$, and let
$$\{Q\}:{S}_{U,k}(1_\lambda,{\varphi})\rightarrow {S}_{\{1\},k_1}(1_\lambda,{\varphi})$$
be as in Lemma \ref{schwspl}.
We define a meromorphic family of maps
$$Q:B_{U,k}(1_\lambda,\tilde{{\varphi}})\rightarrow B_{\{1\},k_1}(1_\lambda,\tilde{{\varphi}})$$
by
$$Q(h):= \{Q\}\left(h - \pi^U_*\circ L\circ [Q]\circ AS(h)\right) + L\circ [Q]\circ AS(h)$$
Then one easily checks that $\pi^U_*\circ Q$ is just the inclusion
$B_{U,k}(1_\lambda,\tilde{{\varphi}})\rightarrow B_{U,k_1}(1_\lambda,\tilde{{\varphi}})$.
\begin{ddd}\label{neuj707}
We define the meromorphic family of restriction maps
$$res^U: D_{\{1\},k_1}(1_\lambda,{\varphi})\rightarrow D_{U,k}(1_\lambda,{\varphi})$$
as the adjoint of $Q$.
\end{ddd}
Note that $res^U$ depends on choices (we do not indicate these choices in the notation for the
restriction map).
In particular, these maps
are not compatible if we change $k$ and $k_1$.
\subsubsection{}\label{neuj2001}
Note that
the composition $res^U\circ ext^U$ coincides with the inclusion
$$D_{U,k_1}(1_\lambda,{\varphi}) \hookrightarrow D_{U,k}(1_\lambda,{\varphi})\ .$$
Recall the Definition \ref{neuj108} of $Ext_{U,k_1}(1_\lambda,{\varphi})$.
\begin{lem}\label{weedef}
Let $k,k_1\in{\mathbb {N}}_0$ and $\lambda\in{\aaaa_\C^\ast}$ be such that
$res^U:D_{\{1\},k_1}(1_\lambda,{\varphi})\rightarrow D_{U,k}(1_\lambda,{\varphi})$ is defined.
Furthermore we assume that
$\pi^U_*:B_{\{1\}}(1_{-\lambda},\tilde{\varphi})\rightarrow
B_U(1_{-\lambda},\tilde{\varphi})$ is regular. Then the restriction
of $res^U$ to $Ext_{U,k_1}(1_\lambda,{\varphi})$ is independent of choices.
\end{lem}
{\it Proof.$\:\:\:\:$} This follows from $res^U\circ ext^U(f)=f$ and the fact that
$ext^U$ is regular at $\lambda$.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}\label{neuj300}
Assume that $\lambda\in{\aaaa_\C^\ast}$ is such that
$$\pi^U_*:B(1_{-\lambda},\tilde{\varphi})\rightarrow
B_U(1_{-\lambda},\tilde{\varphi})$$ and $$[Q]:R_{U}(1_{-\lambda},\tilde{{\varphi}})^n\to R_{\{1\}}(1_{-\lambda},\tilde {\varphi})^n$$ are regular for all $n\in {\mathbb {N}}_0$.
We call $\lambda\in{\aaaa_\C^\ast}$ satisfying these conditions admissible.
\subsubsection{}\label{neuj704}
Note that ${R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^*$
is the space of those distribution sections of $V(1_\lambda,{\varphi})$
that are supported at the point $\infty_P$.
We define $$Ext_{\infty_P}(1_\lambda,{\varphi}):={R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^*\cap Ext_U(1_\lambda,{\varphi})\ .$$
Recall the definition \ref{neuj471} of the space
$Q_{U}(1_\lambda,{\varphi})$. The point of the following lemma is
that every element of $Q_{U}(1_\lambda,{\varphi})$ can be represented by
an invariant distribution supported in $\infty_P$.
Note that $$[ext^{U}]:{R}_{U}(1_{-\lambda},\tilde{\varphi})^*\rightarrow {}^U {R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^*$$
generates a subspace of $Ext_{\infty_P}(1_\lambda,{\varphi})$, where $[ext^{U}]$ is the restriction of $ext^{U}$
to ${R}_{U}(1_{-\lambda},\tilde{\varphi})^*$, or equivalently, the adjoint
of $[\pi^{U}_*]$ (see Def.~\ref{weih254}).
\begin{lem}\label{homjk}
\begin{enumerate}
\item
The family of maps $[ext^{U}]$ generates all of $Ext_{\infty_P}(1_\lambda,{\varphi})$.
\item
If $\lambda\in{\aaaa_\C^\ast}$ is admissible in the sense of \ref{neuj300}, then
there is an exact sequence
$$0\rightarrow
Ext_{\infty_P}(1_\lambda,{\varphi})\rightarrow
{}^{U}{R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^* \rightarrow
Q_{U}(1_\lambda,{\varphi})\rightarrow 0$$
of semisimple $A$-modules.\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
Recall that $${}^{U}{R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^*=
\bigoplus_{n\in{\mathbb {N}}_0} {}^U({R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^n)^*$$
is a weight-decomposition of the $A$-module.
Moreover, we have
$$[ext^{U}]:({R}_{U}(1_{-\lambda},\tilde{\varphi})^n)^*\rightarrow ({R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^n)^*\ .$$
If $f\in Ext_{\infty_P}(1_\lambda,{\varphi})$
is represented by $ext^{U}(h)$ for some meromorphic family
$h_\mu\in D_{U}(1_\mu,{\varphi})$, then
$\{res^{U}\} \{ext^{U}(h)\} =\{h\}$ vanishes at $\lambda$,
where $\{h\}$ denotes the restriction
of $h$ to ${S}_{\{1\}}(1_{-\lambda},\tilde{\varphi})$, and $\{res^U\}:=\{Q\}^*$.
Therefore all non-positive Laurent-coefficients of
the expansion of $h$ at $\lambda$
belong to ${R}_{U}(1_{-\lambda},\tilde{\varphi})^*$.
Thus we can choose the family $h$ such that
$h_\mu\in{R}_{U}(1_{-\mu},\tilde{\varphi})^*$, and such that $[ext^{U}](h)_\lambda=f$.
This shows that $[ext^{U}]$ generates $Ext_{\infty_P}(1_\lambda,{\varphi})$,
and that $A$ acts semisimply on
$Ext_{\infty_P}(1_\lambda,{\varphi})$.
\subsubsection{}
It remains to show that any element of $Q_U(1_\lambda,{\varphi})$
can be represented by some element of ${}^{U}{R}_{\{1\}}(1_{-\lambda},\tilde{\varphi})^*$.
Let $f\in {}^{U}D_{\{1\}}(1_\lambda,{\varphi})$.
Then there is $k\in{\mathbb {N}}_0$ such that $f\in {}^{U}D_{\{1\},k}(1_\lambda,{\varphi})$.
We choose $k_1$ such that $$\{res^{U}\}:{S}_{\{1\},k}(1_{-\lambda},\tilde{\varphi})^*
\rightarrow {S}_{U,k_1}(1_{-\lambda},\tilde{\varphi})^*$$ is defined.
We then put $h:=\{res^{U}\}\{f\}$. Let $$T:{S}_{U,k_1}(1_{-\lambda},\tilde{\varphi})^*
\rightarrow D_{U}(1_\lambda,{\varphi})$$ be the split
induced by the dual split $L$. We
form $f-ext^{U}\circ T(h)$. This difference represents the same
element in $Q_{U}(1_\lambda,{\varphi})$ as $f$, but its restriction to
${S}_{\{1\}}(1_{-\lambda},\tilde{\varphi})$ vanishes. Indeed, for $g$ in ${S}_{\{1\}}(1_{-\lambda},\tilde{\varphi})$
we have
$$ \langle ext^{U}\circ T(h),g\rangle= \langle f, \{Q\}\circ\{\pi_*^U\}(g)\rangle=\langle f, g\rangle
\ .$$ \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\newcommand{{\mathrm{pr}}}{{\mathrm{pr}}}
\newcommand{{\mathbb{F}}}{{\mathbb{F}}}
\subsubsection{}
\begin{prop}\label{argumentprop}
Let $n\ge 0$. If $\lambda\in{\mathfrak{a}}^*$ is sufficiently large, then the inclusion
$$Ext_{\infty_P}(1_\lambda,{\varphi})^n\hookrightarrow {}^U(R_{\{1\}}(1_{-\lambda},\tilde{\varphi})^n)^*$$ is an isomorphism.
\end{prop}
{\it Proof.$\:\:\:\:$}
If $U$ defines a cusp of full rank, then the proposition is a direct consequence of Lemma \ref{vermaolbrich}
and the definition of $[ext^U]$ (Definition \ref{rolf}). Thus we may and will assume in the following
that $U$ defines a cusp of smaller rank.
Note that both spaces appearing in the proposition are finite-dimensional.
The proposition is an immediate consequence of the following lemma.
\begin{lem}\label{argumentlem1}
If $\lambda\in{\mathfrak{a}}^*$ is sufficiently large, then we have the inequality
$$\dim Ext_{\infty_P}(1_\lambda,{\varphi})^n\ge \dim {}^U(R_{\{1\}}(1_{-\lambda},\tilde{\varphi})^n)^*\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
By Lemma \ref{vermaolbrich},2, if $\lambda\in{\mathfrak{a}}^*$ is sufficiently large we have
$\dim {}^U(R_{\{1\}}(1_{-\lambda},\tilde{\varphi})^n)^*=\dim {}^{P_U} {\rm Pol}(N,V_{1_{-\lambda}}\otimes V_{\varphi})^n$, where
the action of $AP_U$ on a polynomial is given by
$$(p \: f)(n):={\varphi}(p)f(n^{p^{-1}})\ .$$
The degree-$n$ subspace of the polynomial maps is characterized by
\begin{equation}\label{argumentdegree1}a\:f=a^{-n\alpha} f\ ,\quad a\in A\ .\end{equation}
We now consider
$$[\pi^U_*]:R_1(1_{-\lambda},\tilde {\varphi})^n\to R_U(1_{-\lambda},\tilde {\varphi})^n$$
which is regular for ${\rm Re }(\lambda)$ sufficiently large. Its adjoint is
$$[ext^U]:(R_U(1_{-\lambda},\tilde {\varphi})^*)^n\to Ext_{\infty_P}(1_\lambda,{\varphi})^n\ ,$$
and we have
$\dim {\mbox{\rm im}} [\pi^U_*] =\dim {\mbox{\rm im}} [ext^U]$.
\subsubsection{}
Therefore,
Lemma \ref{argumentlem1} follows directly from the following lemma.
\begin{lem}\label{argumentlem2}
If $\lambda\in{\mathfrak{a}}^*$ is sufficiently large, then we have
$$\dim {\mbox{\rm im}} [\pi^U_*] \ge \dim {}^{P_U} {\rm Pol}(N,V_{1_{-\lambda}}\otimes V_{\varphi})^n\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
Let $\langle.,.\rangle$ be some non-degenerate invariant bilinear form on ${\mathfrak{g}}$.
We define the linear subspace
\begin{equation}\label{argumenteq33}\bar {\mathfrak{n}}^U:=\{Y\in \bar {\mathfrak{n}}\:|\: \langle [Y,H],X\rangle=0\:\:\forall X\in {\mathfrak{n}}_V\}\end{equation}
and the submanifold $\bar N^U:=\exp(\bar {\mathfrak{n}}^U)\subset \bar N$.
We furthermore choose a Cartan involution $\theta$ of ${\mathfrak{g}}$ compatible with ${\mathfrak{a}}$.
\begin{lem}\label{argumentlem6}
\begin{enumerate}
\item The submanifold $\bar N^U$ is $AM_U$-invariant.
\item The multiplication map
$\bar N_V\times \bar N^U\to \bar N$ is a diffeomorphism, where $\bar N_V:=N_V^\theta$.
\item
The composition
$$\bar N\stackrel{\sim}{\to}\bar N_V\times \bar N^U\stackrel{{\mathrm{pr}}}{\to} \bar N^U$$
is an $AM_U$-equivariant polynomial map.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
The subspace $\bar {\mathfrak{n}}^U$ is $AM_U$-invariant since ${\mathfrak{n}}_V$ is $AM_U$-invariant.
The first assertion now follows from the $AM_U$-equivariance of the exponential map
By the $A$-invariance of $\bar N^U$, $\bar N_V$ and the equivariance of the multiplication, it suffices to show that the multiplication map is a diffeomorphism near $(1,1)$. Infinitesimally it is given by the map $\bar{\mathfrak{n}}_V\times \bar {\mathfrak{n}}^U\to \bar {\mathfrak{n}}$, $(Y_V,Y^U)\mapsto \bar Y_V+Y^U$. This map is an isomorphism. In fact, the dimensions of the domain and the target coincide, and $\bar {\mathfrak{n}}_V\cap \bar {\mathfrak{n}}^U=\{0\}$.
The last assertion follows from the diagram
$$\xymatrix{\bar {\mathfrak{n}}\ar[d]^{\exp}&\bar {\mathfrak{n}}_V\times\bar {\mathfrak{n}}^U\ar[l]_{q\quad}\ar[d]^{(\exp,\exp)}\ar[r]^{\quad pr}&\bar {\mathfrak{n}}^U\ar[d]^{\exp}\\\bar N&\bar N_V\times \bar N^U\ar[l]_{mult\quad}\ar[r]^{\quad{\mathrm{pr}}}&\bar N^U}$$
and the fact that $q$ is an $AM_U$-equivariant polynomial map with a polynomial inverse.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Note that $\bar N_VAM_U$ acts on ${\rm Pol}(\bar N,V_{1_{-\lambda}}\otimes V_{\tilde{\varphi}})$
by
$$\bar n_Vam\: f(\bar n)=\tilde{\varphi}(am) f((\bar n_V^{-1}\bar n)^{(am)^{-1}})\ .$$
We now consider the space
$$I(\lambda)^n:={}^{\bar N_VM_U}{\rm Pol}(\bar N,V_{1_{-\lambda}}\otimes V_{ \tilde {\varphi}})^n\subset{\rm Pol}(\bar N,V_{1_{-\lambda}}\otimes V_{\tilde {\varphi}})^n\ ,$$
where the degree $n$-subspace is distinguished by the condition
\begin{equation}\label{argumentdegree2}a\: f=a^{n\alpha} f\ .\end{equation}
We have a degree-preserving inclusion
$$I(\lambda)^n\hookrightarrow R_1(1_{-\lambda},\tilde {\varphi})^n\ .$$
Furthermore, it follows from Lemma \ref{argumentlem6} that the restriction to $\bar N^U$ induces an isomorphism
$$I(\lambda)^n\stackrel{\sim}{\to} {}^{M_U}{\rm Pol}(\bar N^U,1_{-\lambda}\otimes V_{\tilde {\varphi}})^n\ .$$
Lemma \ref{argumentlem2} now follows from the following two assertions.
\begin{lem}\label{argumentlem3}
$\dim I(\lambda)^n= \dim {}^{P_U} {\rm Pol}(N,V_{1_{-\lambda}}\otimes V_{\varphi})^n$
\end{lem}
\begin{lem}\label{argumentlem4}
If $\lambda\in{\mathfrak{a}}^*$ is sufficiently large, then the restriction of
$[\pi^U_*]$ to $I(\lambda)^n$ is injective.
\end{lem}
\subsubsection{}
We first show Lemma \ref{argumentlem3}.
We define $N^U:=(\bar N^U)^\theta\subset N$. Then we have an $A$-equivariant diffeomorphism
$N_V\times N^U\stackrel{mult}{\to} N$, and the projection $N\stackrel{\sim}{\to} N_V\times N^U\stackrel{{\mathrm{pr}}}{\to} N^U$
is an $AM_U$-equivariant polynomial map. These facts follow from Lemma \ref{argumentlem6} by an application of the Cartan involution $\theta$. We conclude that the restriction to $N^U$ induces an isomorphism
$${}^{P_U} {\rm Pol}(N,V_{1_{-\lambda}}\otimes V_{\varphi})^n\stackrel{\sim}{\to} {}^{M_U}{\rm Pol}(N^U, V_{1_{-\lambda}}\otimes V_{\varphi})^n\ .$$
Note that there is a canonical $AM_U$-equivariant isomorphism between $S(({\mathfrak{n}}^U)^*)\otimes V_{\varphi}$ and \linebreak[4]
${\rm Pol}(N^U, V_{1_{-\lambda}}\otimes V_{\varphi})$. Here
$S(.)$ stands for the symmetric algebra. Similarly, we have
$$S((\bar{\mathfrak{n}}^U)^*)\otimes V_{\tilde{\varphi}}\cong {\rm Pol}(\bar N^U,V_{1_{-\lambda}}\otimes V_{\tilde {\varphi}})\ .$$
The natural pairing between $\bar{\mathfrak{n}}^U$ and ${\mathfrak{n}}^U:= (\bar{\mathfrak{n}}^U)^\theta$ via the $G$-invariant form $\langle.,.\rangle$ induces a nondegenerate pairing between the two symmetric algebras above. We conclude that the spaces
${}^{M_U}{\rm Pol}(N^U, V_{1_{-\lambda}}\otimes V_{\varphi})^n$
and
$
{}^{M_U}{\rm Pol}(\bar N^U,V_{1_{-\lambda}}\otimes V_{\tilde {\varphi}})^n\cong I(\lambda)^n$
are each others duals
(the pairing is degree-preserving in view of the characterizations (\ref{argumentdegree1}) and (\ref{argumentdegree2})).
Lemma \ref{argumentlem3} now follows.
\subsubsection{}
We now start with the proof of Lemma \ref{argumentlem4}.
First of all note that the push-down (see Def.~\ref{weih254})
$$[\pi^U_*]:I(\lambda)^n\to R_U(1_{-\lambda},\tilde {\varphi})^n\subset A_U(1_{-\lambda},\tilde {\varphi})^n $$ is given by a convergent integral
$$[\pi^U_*](f)(\bar n)=\int_{N_V}{\varphi}(n_V)^{-1} f(\bar n(n_V\bar n))a(n_V\bar n)^{-\lambda-\rho} dn_V\ ,\bar n\in \bar N\setminus\{1\}\ .$$
Here we have employed again the Bruhat decomposition $g=\bar n(g) m(g) a(g) n(g)$.
The vector space $V_{\tilde {\varphi}}$ has a filtration induced by the action of $A$ that is preserved by $M_UN_V$. The induced action of $N_V$ on the associated graded vector space
${\mathrm{Gr}}(V_{\tilde {\varphi}})$ is trivial. The filtration of $V_{\tilde {\varphi}}$ induces filtrations on $I(\lambda)^n$ and $A_U(1_{-\lambda},\tilde {\varphi})^n$. From the integral representation of $[\pi^U_*]$ we see that that this map preserves the filtrations and induces a
map ${\mathrm{Gr}}[\pi_U^*]:{\mathrm{Gr}} I(\lambda)^n\to {\mathrm{Gr}} A_U(1_{-\lambda},\tilde {\varphi})^n$.
Since injectivity of the associated graded map implies injectivity of a filtration preserving map, Lemma
\ref{argumentlem4} is a consequence of
\begin{lem}\label{argumentlem44}
If $\lambda\in{\mathfrak{a}}^*$ is sufficiently large, then
$${\mathrm{Gr}} [\pi^U_*]:{\mathrm{Gr}} I(\lambda)^n\to {\mathrm{Gr}} A_U(1_{-\lambda},\tilde {\varphi})^n$$ is injective.
\end{lem}
\subsubsection{}
Note that
$${\mathrm{Gr}} [\pi^U_*](f)(\bar n)=\int_{N_V} f(\bar n(n_V\bar n))a(n_V\bar n)^{-\lambda-\rho}dn_V\ ,\bar n\in \bar N\setminus\{1\}\ .$$
We define
$$c_\lambda(\bar n):={\mathrm{Gr}} [\pi^U_*](1)(\bar n)= \int_{N_V} a(n_V\bar n)^{-\lambda-\rho}dn_V\ ,\bar n\in \bar N\setminus\{1\}\ .$$
We will show the following lemma.
\begin{lem}\label{argumentlem8}
For $\bar n^U\in \bar N^U\setminus\exp(\bar{\mathfrak{n}}_{-2\alpha})$ we have
$$\frac{1}{c_\lambda(\bar n^U)} {\mathrm{Gr}} [\pi^U_*](f)(\bar n^U)= f(\bar n^U)+O(|\lambda|^{-1})\ .$$
\end{lem}
\subsubsection{}
Let us first show that Lemma \ref{argumentlem8} implies \ref{argumentlem44}.
The natural identification $V_{1_{-\lambda}}\cong {\mathbb {C}}$ induces identifications
${\rm Pol}(\bar N^U,1_{-\lambda}\otimes V_{\tilde {\varphi}})^n\cong {\rm Pol}(\bar N^U,V_{\tilde {\varphi}})^n$ for all $\lambda$.
Note that $\bar {\mathfrak{n}}^U\cap\bar{\mathfrak{n}}_{-\alpha}$ is non-trivial. Otherwise we would have
${\mathfrak{n}}_V\cap{\mathfrak{n}}_{\alpha}={\mathfrak{n}}_{\alpha}$ and therefore ${\mathfrak{n}}_V={\mathfrak{n}}$, i.e., $U$ would define a cusp of full rank.
Hence we can choose a finite sequence of base points $\bar n^U_i\in \bar N^U\setminus \exp(\bar{\mathfrak{n}}_{-2\alpha})$ and vectors
$v_i\in V_{\varphi}$, $i=1,\dots,r:=\dim {}^{M_U}{\rm Pol}(\bar N^U,V_{\tilde {\varphi}})^n$ such that
the following map is an isomorphism:
$$\Phi:{}^{M_U}{\rm Pol}(\bar N^U,V_{\tilde {\varphi}})^n\to{\mathbb {C}}^r\ , f\mapsto (\langle v_1,f(n^U_1)\rangle,\dots,\langle v_r,f(n^U_r)\rangle)\ .$$
We now consider the composition
$$A(\lambda):=\Phi\circ \frac{1}{c_\lambda} {\mathrm{Gr}} [\pi^U_*]\circ \Phi^{-1}:{\mathbb {C}}^n\to {\mathbb {C}}^n\ .$$
Lemma \ref{argumentlem8} implies that
$$A(\lambda)=1+O(|\lambda|^{-1})\ .$$
In particular, $A$ is injective, if $\lambda\in{\mathfrak{a}}$ is sufficiently large.
This implies the assertion of Lemma \ref{argumentlem44}.
\subsubsection{}
We now show Lemma \ref{argumentlem8}.
Fix $\bar n^U\in \bar N^U\setminus \exp(\bar{\mathfrak{n}}_{-2\alpha})$. We set
$$\Psi(n_V):=\log a(n_V\bar n^U) \ ,\quad g(n_V):=f(\bar n(n_V\bar n^U))$$
Then we can write
$$\int_{N_V} f(\bar n(n_V\bar n^U))a(n_V\bar n^U)^{-\lambda-\rho}dn_V=\int_{N_V} {\rm e}^{(-\lambda-\rho)\Psi( n_V)}g(n_V)dn_V\ .$$
\begin{lem}\label{argumentlem9}
The function $\Psi(n_V)$ has a unique non-degenerate absolute minimum $\Psi(1)=0$ at
$n_V=1$. Furthermore, there exists a compact neighbourhood $K\subset N_V$ of $1$ such that
$\alpha(\Psi(n_V))\ge 1$ for $n_V\not\in K$.
\end{lem}
We first show that Lemma \ref{argumentlem9} implies \ref{argumentlem8}.
We split the integral
as $\int_{N_V}=\int_{K}+\int_{N_V\setminus K}$.
Note that $\bar n(N_V\bar n^U)\subset \bar N$ is a pre-compact subset. It follows that
$g$ is smooth and uniformly bounded. We approximate
the first summand by a Gaussian integral at the minimum of $\Psi$ and get
$$\int_K {\rm e}^{(-\lambda-\rho)(\Psi(n_V))}g(n_V)dn_V= \int_K {\rm e}^{(-\lambda-\rho)(\Psi(n_V))}dn_V \left(g(0)+O(\lambda^{-1})\right)\ .$$
Furthermore,
$\int_K {\rm e}^{(-\lambda-\rho)(\Psi(n_V))}dn_V$ decreases at most as $|\lambda|^{-\dim N_V/2}$.
We claim that the contributions $\int_{N_V\setminus K}$ decrease exponentially so that these parts of the integrals
can only contribute exponentially small error terms. This claim implies Lemma \ref{argumentlem8}.
In order to see the claim
we write
$$\int_{N_V\setminus K}{\rm e}^{(-\lambda-\rho)(\Psi(n_V))}dn_V =\min_{n_V\in N_V\setminus K}{\rm e}^{(-\lambda+\alpha)(\Psi(n_V))}\int_{N_V\setminus K}{\rm e}^{(-\alpha-\rho)(\Psi(n_V))}dn_V\ .$$
The integral on the right-hand side converges, and
$$\min_{n_V\in N_V\setminus K}{\rm e}^{(-\lambda+\alpha)(\Psi(n_V))}\le {\rm e}^{-c|\lambda|}$$ for a suitable constant $c>0$.
This finishes the proof of Lemma \ref{argumentlem8} under the assumption of Lemma \ref{argumentlem9}.
\subsubsection{}
We now show Lemma \ref{argumentlem9}. In order to compute $a(n_V\bar n^U)$ we may assume
that $G$ is the subgroup of $GL(n+1,{\mathbb{F}})$ that preserves the ${\mathbb{F}}$-valued Hermitian scalar product
\begin{equation}\label{argumenteq45}\bar v_0 w_n+ \bar v_n w_0+\bar v_1 w_1+ \dots + \bar v_{n-1} w_{n-1}\ \end{equation}
on the right ${\mathbb{F}}$-vector space ${\mathbb{F}}^{n+1}$.
We choose
$$A:=\left\{\left(\begin{array}{ccc}a&0&0\\0&1_{n-1\times n-1}&0\\0&0&a^{-1}\end{array}\right)\:|\: a\in {\mathbb{R}}^+\right\}$$
and get
$$N:=\left\{\left(\begin{array}{ccc}1&w&p-\frac{\|w\|^2}{2}\\0&1&-\bar w^t\\0&0&1\end{array}\right)\:|\: w\in {\mathbb{F}}^{n-1}\ ,p\in {\mbox{\rm im}}\, {\mathbb{F}}\right\}\ .$$
The parametrization of $N$ given here is via the exponential map, if we identify
${\mathfrak{n}}={\mathfrak{n}}_\alpha\oplus{\mathfrak{n}}_{2\alpha}\cong {\mathbb{F}}^{n-1}\oplus {\rm Im}\,{\mathbb{F}}$.
Furthermore,
$$\bar N:=\left\{\left(\begin{array}{ccc}1&0&0\\-\bar v^t&1&0\\q-\frac{\|v\|^2}{2}&v&1\end{array}\right)\:|\: v\in {\mathbb{F}}^{n-1}\ ,q\in {\rm Im}\, {\mathbb{F}}\right\}\ .$$
Since they are $A$-invariant, the subspaces ${\mathfrak{n}}_V\subset {\mathfrak{n}}$ and $\bar {\mathfrak{n}}^U\subset \bar {\mathfrak{n}}$ are given in this identification as
$W\oplus P\subset {\mathfrak{n}}$ and $V\oplus Q\subset \bar {\mathfrak{n}}$ for real subspaces $W,V\subset {\mathbb{F}}^{n-1}$ and $P,Q\subset {\rm Im}\, {\mathbb{F}}$.
We consider the invariant form $\langle A,B\rangle:={\rm Re }\,{\rm Tr} AB$ on ${\mathfrak{g}}$. Given
$W\oplus P$, the space $V\oplus Q$ is characterized by
$\langle{\mbox{\rm ad}}(V\oplus W)(H),W\oplus H\rangle=0$ (see (\ref{argumenteq33})).
Explicitly,
$$H=\left(\begin{array}{ccc}1&0&0\\0&0&0\\0&0&-1\end{array}\right)\ ,\langle{\mbox{\rm ad}}(v+q)(H),w+p\rangle=-2{\rm Re } (\bar w v^t+qp)\ .$$
Therefore we have
\begin{equation}\label{argumenteq88}V=W^\perp\ , Q=P^\perp\end{equation} with respect to the natural euclidean pairings on ${\mathbb{F}}^{n-1}$ and ${\rm Im}\, {\mathbb{F}}$.
The vector $e_0:=(1,\dots,0)\in {\mathbb{F}}^{n+1}$ is the highest weight vector of the standard representation of $G$ with weight $\alpha$. Similarly, $e_n:=(0,\dots,0,1)$ is the lowest weight vector with weight $-\alpha$.
Let $g=\bar n man\in \bar NMAN$. If $\langle.,.\rangle$ denotes the ${\mathbb{F}}$-valued scalar product (\ref{argumenteq45}), then
$$\langle ge_0,e_n\rangle=\langle\bar ma n e_0,e_n\rangle=\langle ma n e_0,\bar n^{-1}e_n\rangle=a^\alpha\langle m e_0,e_n\rangle\ .$$
The subspace $e_0{\mathbb{F}}\cong {\mathbb{F}}$ is invariant under the group $M$. In particular, we have a homomorphism
$\vartheta:M\to {\mathbb{F}}^*$ such that $m e_0=e_0 \vartheta(m)$. It now follows that
$\langle m e_0,e_n\rangle=\bar \vartheta(m)$. Since $M$ is compact, we have $\|\vartheta(m)\|=1$ and therefore
$a^\alpha=\|\langle ge_0,e_n\rangle\|$.
If we parametrize $(w,p)=n_V\in N_V$ and $(v,q)=\bar n^U\in \bar N^U$ as above, then we get
\begin{equation}\label{argumenteq100} a(n_Vn^U)^{2\alpha}=\|1-w\bar v^t+(p-\frac{\|w\|^2}{2})(q-\frac{\|v\|^2}{2})\|^2\ .\end{equation}
We fix $(v,q)\in V\oplus Q$ with $v\not=0$.
We must show that the right-hand side has a unique absolute minimum $1$ at $(w,p)=(0,0)$, and that this minimum is non-degenerate. Using (\ref{argumenteq88}) we get
\begin{eqnarray*}
{\rm Re }(1-w\bar v^t+(p-\frac{\|w\|^2}{2})(q-\frac{\|v\|^2}{2}))&=&1+\frac{\|v\|^2\|w\|^2}{4}\\
{\rm Im}(1-w\bar v^t+(p-\frac{\|w\|^2}{2})(q-\frac{\|v\|^2}{2}))&=&pq-w\bar v^t-\frac{\|w\|^2}{2}q-\frac{\|v\|^2}{2}p\ .
\end{eqnarray*}
First of all, $a(n_Vn^U)^{2\alpha}\ge 1$ and $a(1)^{2\alpha}=1$. Moreover, if
$a(n_V\bar n^U)^{2\alpha}=1$, then $w=0$.
Since ${\rm Re }(\overline{pq}p)={\rm Re } \bar q\|p\|^2)=0$ we have
$pq\perp \frac{\|v\|^2}{2}p$. Hence, the equality $a(n_V\bar n^U)=1$ implies in addition to $w=0$ that also $p=0$.
We thus have shown that $a(n_V\bar n^U)^{2\alpha}$ takes its unique absolute minimum at $1$.
Next we show that it is non-degenerate. The Hessian $h(p,w)$ is the part of the polynomial (\ref{argumenteq100}) which is quadratic in $(w,p)$. It can be written as
$$h(w,p):=\|w\|^2\frac{\|v\|^2}{2}+\|p(q-\frac{\|v\|^2}{2})-w\bar v^t\|^2\ .$$
If $h(w,p)=0$, then from the first summand and $v\not=0$ we get $w=0$ and $p(q-\frac{\|v\|^2}{2})=0$. Since $(q-\frac{\|v\|^2}{2})\not=0$ we conclude that $p=0$.
The last assertion of Lemma \ref{argumentlem9} follows from
$\lim_{n_V\to \infty}a(n_V\bar n^U)^{2\alpha}=\infty$, which is easy to check using the explicit formula (\ref{argumenteq100}). This finishes the proof of Lemma \ref{argumentlem9}.
We now have also finished the proof of Proposition \ref{argumentprop}.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
\begin{prop}\label{maertins}
\begin{enumerate}
\item We have $Ext_{U}(1_\lambda,{\varphi})=E_{U}(1_\lambda,{\varphi})$
for all $\lambda\in{\aaaa_\C^\ast}$.
\item If $\lambda\not\in I_{\mathfrak{a}}$ or ${\rm Re }(\lambda)>-\rho$, then
$Q_U(1_\lambda,{\varphi})=0$.\end{enumerate}
\end{prop}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
As in\ref{neuj703} let ${\mathcal{E}}_{\infty_P}(1,{\varphi})^n$ be the torsion-free
coherent sheaf on ${\aaaa_\C^\ast}$ of holomorphic families of
$U$-invariant distributions supported on $\infty_P$,
on which $A$ acts by multiplication by the function
$\lambda\mapsto a^{\lambda-\rho-n\alpha}$.
By ${\mathcal{E}} xt_{\infty_P}(1,{\varphi})^n$ we denote the subsheaf generated by the
restriction of $[ext^{U}]$ to the homogeneous part $({R}_{U}(1_{-.},\tilde{\varphi})^n)^*$.
The space $Ext_{\infty_P}(1_\lambda,{\varphi})^n$ is the geometric fibre of ${\mathcal{E}} xt_{\infty_P}(1,{\varphi})^n$ for the generic set of $\lambda\in{\aaaa_\C^\ast}$, where $[ext^U]$ is regular.
\subsubsection{}\label{neuj705}
The sheaf ${\mathcal{E}}_{\infty_P}(1,{\varphi})^n$ is the sheaf of sections of a finite-dimensional trivial holomorphic vector bundle $E_{\infty_P}(1,{\varphi})^n\to {\aaaa_\C^\ast}$.
The torsion-free subsheaf ${\mathcal{E}} xt_{\infty_P}(1,{\varphi})^n$ corresponds to a
bundle ${\mathrm{ Ext}}_{\infty_P}(1,{\varphi})^n$.
This discussion shows the following: if the inclusion $Ext_{\infty_P}(1_\lambda,{\varphi})^n\hookrightarrow E_{\infty_P}(1_\lambda,{\varphi})^n$ is surjective
at one point $\lambda\in {\aaaa_\C^\ast}$, then it is surjective for generic $\lambda$, i.e. outside a discrete set.
However, we know from Proposition \ref{argumentprop} that this inclusion is surjective for many $\lambda$, hence it is so generically.
\subsubsection{}\label{miau}
In view of Lemma \ref{homjk}, 2.
we have
\begin{equation}\label{neuj708}
E_{\infty_P}(1_\lambda,{\varphi})^n/E xt_{\infty_P}(1_\lambda,{\varphi})^n\cong \ker\left( Q_U(1_\lambda,{\varphi})^n\to
\bar Q_{\infty_P}(1_\lambda,{\varphi})^n\right)\ .
\end{equation}
By \ref{neuj705} the quotient on the left hand side is trivial generically. The same is true
for $\bar Q_{\infty_P}(1_\lambda,{\varphi})^n$ by Lemma \ref{vermaolbrich}. We conclude that $Q_U(1_\lambda,{\varphi})^n$ is trivial outside a discrete set.
\subsubsection{}
We now prove the first assertion of Prop.~\ref{maertins}.
We know that $Ext_{U}(1_\lambda,{\varphi})\subset E_{U}(1_\lambda,{\varphi})$.
Let now $f\in E_{U}(1_\lambda,{\varphi})$ be given
as the value at $\lambda$
of a meromorphic family $f_\mu\in {}^{U} D_{\{1\},k}(1_\mu,{\varphi})$ for some sufficiently large $k$.
Let $res^U$ be the meromorphic family of continuous maps
$res^U:{}^{U} D_{\{1\},k}(1_\mu,{\varphi}):\rightarrow D_{U,k_1}(1_\mu,{\varphi})$
for suitable $k_1\in{\mathbb {N}}$ (see Definition \ref{neuj707}). By \ref{miau} for generic
$\mu$ we can write $f_\mu=ext^{U} g_\mu$ for some $g_\mu\in
D_{U,k}(1_\mu,{\varphi})$.
We conclude $ext^{U}\circ res^{U}(f_\mu)= ext^{U}\circ res^{U} \circ ext^U(g_\mu)= ext^U(g_\mu)=f_\mu$ by Lemma \ref{weedef}.
Thus
\begin{equation}\label{wau}
ext^{U}\circ res^{U}(f_\mu)=f_\mu \quad\mbox{ for all }\mu\ .
\end{equation}
We conclude that $f\in Ext_{U}(1_\lambda,{\varphi})$.
\subsubsection{}
We now turn to the second assertion. The first assertion implies that the left hand side
of (\ref{neuj708}) is trivial for all $\lambda$. Therefore the map $Q_U(1_\lambda,{\varphi})^n\to
\bar Q_{\infty_P}(1_\lambda,{\varphi})^n$ is always injective. We now apply Lemma \ref{vermaolbrich}.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
The argument that proofs Equation (\ref{wau}) also shows the following.
\begin{lem}\label{wauwau}
The composition $ext^U\circ res^U$
is the identity on $Ext_{U}(1_{\lambda},{\varphi})$.
\end{lem}
\subsection{The scattering matrix}
\subsubsection{}
Recall that the family of intertwining operators (see \ref{gaa100} for the unnormalized version and \cite{MR1749869} for normalizations) forms a meromorphic family of operators.
It therefore maps (holomorphic) families of invariant sections to (meromorphic) families.
It follows that
$J_\lambda$ maps $E_U(1_\lambda,{\varphi})$ to $E_U(1_{-\lambda},{\varphi})$
if $\lambda\in {\aaaa_\C^\ast}$ is such that $J_\lambda$ is regular (e.g. $\lambda\not\in I_{\mathfrak{a}}$, see \ref{gaa101}). By Proposition \ref{maertins} we get a mapping
\begin{equation}\label{uhu}
J_\lambda:Ext_{U}(1_\lambda,{\varphi})\rightarrow Ext_{U}(1_{-\lambda},{\varphi})\ .
\end{equation}
Moreover,
for given $k\in{\mathbb {N}}_0$
we have
$$J_\lambda:Ext_{U,k}(1_\lambda,{\varphi})\rightarrow Ext_{U,k_1}(1_{-\lambda},{\varphi})$$
if $k_1\in{\mathbb {N}}_0$ is sufficiently large.
\subsubsection{}\label{neuj2004}
We can now define the scattering matrix
$$S^U_\lambda:D_U(1_\lambda,{\varphi})\rightarrow D_U(1_{-\lambda},{\varphi})\ .$$
Fix $k\in{\mathbb {N}}_0$ and a compact subset $W\subset {\mathfrak{a}}^*$.
Then we choose $k_0>k$ and $k_1,k_2$ such that
$$J_\lambda: Ext_{U,k_0}(1_\lambda,{\varphi})\rightarrow Ext_{U,k_1}(1_{-\lambda},{\varphi})$$
for generic $\lambda$ (e.g. non-integral) with ${\rm Re }(\lambda)\in W$
and
$$res^U:D_{\{1\},k_1}(1_\lambda,{\varphi})\rightarrow D_{U,k_2}(1_\lambda,{\varphi})$$
is defined (Definition \ref{neuj707}) as a meromorphic family on $W$.
Then we consider the composition
$$S^U_\lambda:=res^U\circ J_\lambda\circ ext^U
:D_{U,k}(1_\lambda,{\varphi})\rightarrow D_{U,k_2}(1_{-\lambda},{\varphi})\ .$$
By definition $S^U_\lambda$ is a meromorphic family of continuous maps.
Since
$$J_\lambda\circ ext^U:D_U(1_\lambda,{\varphi})\to Ext_U(1_{-\lambda},{\varphi})$$
Lemma \ref{weedef} now implies that $S^U_\lambda$
is well-defined independently of the choices made for $res^U$.
\subsubsection{}
Letting $k$ tend to infinity and
$W$ run over a sequence of compact subsets exhausting ${\mathfrak{a}}^*$
we are arrive at the following definition.
\begin{ddd}\label{neuj2003}
We define the scattering matrix as the meromorphic family of continuous maps
$$S^U_\lambda:D_U(1_\lambda,{\varphi})\rightarrow D_U(1_{-\lambda},{\varphi})\ ,$$
which is given by the composition
$$S^U_\lambda(f):=res^U\circ J_\lambda\circ ext^U(f)\ $$
whenever the constituents are regular.
\end{ddd}
\subsubsection{}
If $-\lambda$ is admissible in the sense of \ref{neuj300} (with ${\varphi}$ replaced by $\tilde {\varphi}$),
then the push-down
$$\pi^U_*:B_{\{1\}}(1_\lambda,\tilde{{\varphi}})\to B_U(1_\lambda,\tilde{{\varphi}})$$ is regular and admits a right-inverse. Hence it induces an isomorphism
$$B_U(1_\lambda,\tilde{{\varphi}})\cong B_{\{1\}}(1_\lambda,\tilde{{\varphi}})/\ker\pi^U_*\ .$$
If $\lambda\not\in I_{\mathfrak{a}}$, then $J_\lambda$ is regular. We claim that it maps $\ker\pi^U_*$
to the kernel of $\pi^*_U$ at $-\lambda$. Let $f\in\ker\pi^U_*$ and $\phi\in D_U(1_{\lambda},{\varphi})$.
By (\ref{uhu}) there is an $\psi\in\ D_U(1_{-\lambda},{\varphi})$ such that
$J_\lambda\circ ext^U(\phi)=ext^U(\psi)$. We find
$$
\langle \pi^U_*\circ J_\lambda (f),\phi\rangle=
\langle f,J_\lambda\circ ext^U(\phi)=
\langle f,ext^U(\psi)\rangle=
\langle \pi^U_*(f),\psi\rangle=0\ .
$$
The claim follows.
Therefore, if $\pm\lambda$ is admissible and non-integral, then the operator $J_\lambda$
descends to a map
\begin{equation}\label{neuj801}
\tilde S^U_\lambda:
B_U(1_\lambda,\tilde{{\varphi}})\rightarrow B_U(1_{-\lambda},\tilde{{\varphi}})\ .
\end{equation}
On the other hand, we have a meromorphic family of maps given by the adjoint of the scattering matrix
$${}^tS^U_\lambda:B_U(1_\lambda,\tilde{{\varphi}})\rightarrow B_U(1_{-\lambda},\tilde{{\varphi}})\ .$$
\begin{lem} \label{gen}
If $\pm\lambda$ is non-integral and admissible, then we have
$\tilde S^U_\lambda={}^tS^U_\lambda$.
\end{lem}
{\it Proof.$\:\:\:\:$}
Recall from Lemma \ref{wauwau} that $ext^U\circ res^U$
is the identity on $Ext_{U}(1_{-\lambda},{\varphi})$. For non-integral $\lambda$ we have $Ext_{U}(1_{-\lambda},{\varphi})={}^U C^{-\infty}(\partial X,V(1_{-\lambda},{\varphi}))$ by Proposition \ref{maertins},~2.
The assumptions on $\lambda$ imply that
$\tilde S^U_\lambda$ is defined, that
$\pi^U_*:B_{\{1\}}(1_\lambda,\tilde{{\varphi}})\rightarrow B_U(1_\lambda,\tilde{{\varphi}})$, $ext^U: D_U(1_{\lambda},{\varphi})\rightarrow
D_{\{1\}}(1_{\lambda},{\varphi})$, and $J_\lambda$ are regular.
Let $f\in B_U(1_\lambda,\tilde{{\varphi}})$ be given by $\pi^U_*(F)$, $F\in B_{\{1\}}(1_{\lambda},{\varphi})$.
Furthermore let $\phi\in D_U(1_{\lambda},{\varphi})$.
Then we compute
\begin{eqnarray*}
\langle \tilde S^U_\lambda(f),\phi\rangle&=&
\langle \pi^U_*\circ J_\lambda(F),\phi\rangle\\
&=&\langle F,J_\lambda\circ ext^U(\phi)\rangle\\
&=&\langle F,ext^U\circ res^U\circ J_\lambda\circ ext^U(\phi)\rangle\\
&=&\langle f,S^U_\lambda(\phi)\rangle\\
&=&\langle {}^tS^U_\lambda(f),\phi\rangle\ .
\end{eqnarray*}
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Note that the normalization of ${\varphi}$ in the present subsection differs
from that in \ref{weih213}, since we want $\tilde {\varphi}$ to be normalized as required there.
As a consquence
the space $R_{U}(1_\lambda,{\varphi})$ contains summands with negative index.
If we put $k_{\varphi}:=l_{\varphi}/\alpha$, then we
have $$R_{U}(1_\lambda,{\varphi})=\prod^\infty_{n=-k_{\varphi}} R_{U}(1_\lambda,{\varphi})^n\ ,\quad
B_U(1_\lambda,{\varphi})=\bigcap_{k\ge -k_{\varphi}}B_{U,k}(1_\lambda,{\varphi})\ .$$
\subsubsection{}
\begin{lem}\label{fundd}
Assume that $U$ defines a cusp of smaller rank.
If $l_{\varphi}<2\rho^U$, then there is a natural non-degenerate pairing
between
$B_{U,-k_{\varphi}}(1_\lambda,{\varphi})$ and $B_{U,0}(1_{-\lambda},\tilde{\varphi})$
given by integration over $B_U$. We obtain an inclusion
$$B_{U}(1_\lambda,{\varphi})\subset B_{U,-k_{\varphi}}(1_\lambda,{\varphi})\hookrightarrow D_U(1_\lambda,{\varphi})\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
Let $f\in B_{U,-k_{\varphi}}(1_\lambda,{\varphi})$ and $\phi\in B_{U,0}(1_{-\lambda},\tilde{\varphi})$.
Let $W\subset N\setminus N_V$ be any compact subset.
Then there exists a constant $C\in {\mathbb{R}}$ such that for all $\xi\in W$ and
$a\in A_+$ we have
\begin{eqnarray}
|\langle f(\xi^aw),\phi(\xi^aw) \rangle |
&=& |\langle{\varphi}(a){\varphi}(a)^{-1} f(\xi^aw),\phi(\xi^aw) \rangle|\nonumber\\
&=&|\langle {\varphi}(a)^{-1} f(\xi^aw),\tilde{\varphi}(a)^{-1}\phi(\xi^aw)\rangle|\nonumber\\
&\le& C \|f\|_{-k_{\varphi},0} \|\phi\|_{0,0} a^{l_{\varphi}-4\rho^U}\label{stegh}
\end{eqnarray}
(see \ref{weih136} for the definition of the norms).
\subsubsection{}
Let $F\subset N_V$ be any compact fundamental domain for the lattice $V\subset N_V$.
We can now write
\begin{eqnarray*}
&&\int_{U\backslash N}|\langle f(xw),\phi(xw) \rangle | dx
\\&=&\frac{1}{[U:U^0]}\int_{N_V\backslash N}
\int_{F} |\langle f(yvw),\phi(yvw) \rangle |dv dy\\
&=&\frac{1}{[U:U^0]}\int_A \int_{S(N_V\backslash N)}
\int_{F} |\langle f(\xi^a vw),\phi(\xi^avw) \rangle |dv d\xi a^{2\rho^U} da\\
&=&\frac{1}{[U:U^0]}\int_A \int_{S(N_V\backslash N)}
\int_{F} |\langle f((\xi v^{a^{-1}})^aw),\phi((\xi v^{a^{-1}})^aw) \rangle |dv d\xi a^{2\rho^U} da\\
&=& I_+ + I_-\ ,
\end{eqnarray*}
where $I_\pm$ are the integrals over $A_\pm$, respectively.
Inserting (\ref{stegh}) we obtain
$$
\int_{S(N_V\backslash N)}
\int_{F} |\langle f((\xi v^{a^{-1}})^aw),\phi((\xi v^{a^{-1}})^aw) \rangle |dv d\xi \le C_1 \|f\|_{-k_{\varphi},0} \|\phi\|_{0,0}a^{l_{\varphi}-4\rho^U}$$
and therefore
$$I_+ \le C_2 \|f\|_{-k_{\varphi},0} \|\phi\|_{0,0} \int_{ A_+}a^{l_{\varphi}-2\rho^U}da \ .$$
The integral on the right-hand side converges since $l_{\varphi}-2\rho^U<0$
by assumption. Since
$I_-$ can clearly be estimated by $C_3 \|f\|_{-k_{\varphi},0} \|\phi\|_{0,0}$,
we have shown the lemma. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}\label{neuj2005}
Using the inclusion (Lemma \ref{fundd})
$$B_{U}(1_\lambda,{\varphi})\hookrightarrow D_U(1_\lambda,{\varphi})$$
we can consider the restriction of the scattering matrix
$$(S^U_\lambda)_{|B_{U}(1_\lambda,{\varphi})}:B_{U}(1_\lambda,{\varphi})\to D_{U}(1_{-\lambda},{\varphi})\ .$$
If $f\in B_{U}(1_\lambda,{\varphi})$, then
the restriction of the distribution $ext^U(f)$ to $\Omega_P$ is smooth.
Since $J_\lambda$ is pseudo-local it
it follows that
$J_\lambda\circ ext(f)$ is smooth on $\Omega_P$. Hence the restriction of the distribution
$S^U_\lambda(f)$ to the Schwartz space
${S}_U(1_\lambda,\tilde{\varphi})$ is given by integration against a smooth section.
The following lemma asserts that $S^U_\lambda(f)={}^tS_\lambda^U(f)\in B_{U}(1_{\lambda},{\varphi})$ considered as an element of $D_{U}(1_{-\lambda},{\varphi})$ via Lemma \ref{fundd}.
\subsubsection{}
\begin{lem}\label{inside}
Assume that $U$ defines a cusp of smaller rank and that $l_{\varphi} <2\rho^U$.
Then we have an equality of meromorphic families
\begin{equation}\label{neuj900}
(S^U_\lambda)_{|B_{U}(1_\lambda,{\varphi})}={}^tS^U_\lambda\ .
\end{equation}
\end{lem}
{\it Proof.$\:\:\:\:$}
By meromorphic continuation it suffices to show
the equality (\ref{neuj900}) for non-integral $\lambda$ in the
open subset
$\{|{\rm Re }(\lambda)|<\rho^U-l_{\varphi}/2\}\subset {\aaaa_\C^\ast}$ such that $-\lambda$ is admissible in the sense of \ref{neuj300}.
Then the push-down converges at $\pm\lambda$.
Let $f\in B_{\{1\}}(1_\lambda,\tilde{\varphi})$.
We consider an element $\phi\in {S}_{\{1\}}(1_\lambda, {\varphi})$.
We view $\tilde S^U_\lambda\circ \pi^U_*(f)$ (see (\ref{neuj801})) as an element
of $D_U(1_{-\lambda},\tilde{\varphi})$ and compute
\begin{eqnarray*}
\langle \tilde S^U_\lambda\circ \pi^U_*(f),\pi^U_*(\phi)\rangle_{B_U} &=&
\langle \pi^U_*\circ J_\lambda (f),\pi^U_*(\phi)\rangle_{B_U}\\
&=&\langle \pi^U_*\circ J_\lambda (f),\phi\rangle_{\partial X}\\
&=&\sum_{u\in U} \langle \pi^{1_{-\lambda},\tilde{\varphi}}(u) J_\lambda(f),\phi\rangle_{\partial X}\\
&=&\sum_{u\in U} \langle f,\pi^{1_{-\lambda},{\varphi}}(u) J_\lambda(\phi)\rangle_{\partial X}\\
&=&\langle f,\pi^U_* \circ J_\lambda (\phi)\rangle_{\partial X}\\
&=&\langle \pi_*^U(f),\pi^U_* \circ J_\lambda (\phi)\rangle_{B_U}\\
&=&\langle ext^U\circ \pi_*^U(f),J_\lambda (\phi)\rangle_{\partial X}\\
&=&\langle J_\lambda\circ ext^U\circ \pi^U_*(f),\phi\rangle_{\partial X}\\
&=&\langle ext^U\circ res^U\circ J_\lambda\circ ext^U\circ \pi^U_*(f),\phi\rangle_{\partial X}\\
&=&\langle S^U_\lambda\circ \pi^U_*(f),\pi^U_*(\phi)\rangle_{B_U}\ .
\end{eqnarray*}
Since the $\pi^U_*(\phi)$, $\phi\in {S}_{\{1\}}(1_\lambda,{\varphi})$ (resp. $\pi_*^U(f)\in B_{U}(1_\lambda,\tilde{\varphi})$, $f\in B_{\{1\}}(1_\lambda,\tilde{\varphi})$),
exhaust ${S}_U(1_\lambda,{\varphi})$ (resp. $B_{U}(1_\lambda,\tilde{\varphi})$) we conclude that
$$\tilde S^U_\lambda(h) =S^U_\lambda(h) $$ as smooth sections for all $h\in B_{U}(1_\lambda,\tilde{\varphi})$
and hence $${}^tS^U_\lambda = (S^U_\lambda)_{|B_{U}(1_\lambda,{\varphi})} $$
by Lemma \ref{gen}. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Next we show that the scattering matrix is off-diagonally
smoothing. Let $\chi,\tilde\chi$ be smooth cut-off functions
on $B_U$ with compact support such that
$\tilde\chi \chi = \chi$.
\begin{lem}\label{offside}
Assume that $U$ defines a cusp of smaller rank and that $l_{\varphi} <2\rho^U$.
Then we have a meromorphic family
of continuous maps
$$(1-\tilde\chi)\circ S^U_\lambda \circ \chi:D_U(1_\lambda,{\varphi})\rightarrow B_U(1_{-\lambda},{\varphi})\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
We are going employ the fact that $J_\lambda$ is off-diagonally
smoothing.
Let $\chi^U\in C^\infty(\Omega_P)$ be a cut-off function such that
$\sum_{u\in U} u^* \chi^U\equiv 1$, and such that
the restriction
of the projection $\Omega_P\rightarrow B_U$ to ${\mathrm{supp}}(\chi^U)$
is proper. Then there exists a cut-off function
$\bar\chi$ on $\partial X$
such that $\bar\chi \chi\chi^U=0$ and
$\bar\chi(1-\tilde\chi)=1-\tilde\chi$.
Finally we choose a compactly supported cut-off function $\hat\chi$
on $B_U$ such that $\chi^U \bar \chi \hat \chi=0$ and $\hat \chi \chi=\chi$.
Let $\phi\in {S}_U(1_{\lambda},\tilde{\varphi})$ be any test function
and $f\in D_U(1_\lambda,{\varphi})$. We approximate
$\chi f$ by a sequence of smooth functions
$f_\alpha$ with compact support.
We can in addition assume that ${\mathrm{supp}} f_\alpha\subset \{\hat \chi=1\}$
for all $\alpha$.
Then we compute
\begin{eqnarray*}
\langle (1-\tilde\chi)S^U_\lambda(\chi f),\phi\rangle_{B_U}&=&
\lim_{\alpha} \langle (1-\tilde\chi)S^U_\lambda f_\alpha,\phi\rangle_{B_U}\\
&\stackrel{\mathrm{Lemma}\,\ref{inside}}{=}&\lim_{\alpha} \langle (1-\tilde\chi){}^tS^U_\lambda f_\alpha,\phi\rangle_{B_U}\\
&=&\lim_{\alpha} \langle f_\alpha, S^U_\lambda( (1-\tilde\chi) \phi)\rangle_{B_U}\\
&=&\lim_{\alpha} \langle f_\alpha, \{res^U\} \circ J_\lambda \circ ext^U((1-\tilde\chi) \phi)\rangle_{B_U}\\
&=&\lim_{\alpha} \langle \chi^U f_\alpha , J_\lambda \circ ext^U((1-\tilde\chi) \phi)\rangle_{\partial X}\\&=&\lim_{\alpha} \langle \chi^U\hat \chi f_\alpha , J_\lambda \circ \bar \chi ext^U((1-\tilde\chi) \phi)\rangle_{\partial X}\\
&=&\lim_{\alpha}\langle \pi^U_* (\bar\chi J_\lambda (\chi^U \hat \chi f_\alpha)),(1-\tilde\chi) \phi \rangle_{B_U}\\
&\stackrel{(*)}{=}&\langle \pi^U_* \left(\bar\chi J_\lambda ( \chi^U \chi f)-\tilde\chi\pi^U_* (\bar\chi J_\lambda (\chi \chi^U f))\right) , \phi \rangle_{B_U}\ .
\end{eqnarray*}
In order to see the equality marked by $(*)$ note that
that $\bar\chi J_\lambda \hat\chi\chi^U$
is a continuous map from $D_U(1_\lambda,{\varphi})$ to $B_{\{1\}}(1_{-\lambda},{\varphi})$
(since $J_\lambda$ is off-diagonal smoothing) and that
$\lim_{\alpha}\chi^U \hat \chi f_\alpha=\chi^U\chi f$. The assertion of the Lemma follows from
the identity proved above:
$$(1-\tilde\chi)\circ S^U_\lambda\circ \chi (.) = \pi^U_* \left(\bar\chi J_\lambda (\chi \chi^U .)-\tilde\chi \pi^U_* (\bar\chi J_\lambda (\chi \chi^U .))\right)\ .$$
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\section{The general case}
\subsection{The space $B_\Gamma(\sigma_\lambda,{\varphi})$}\label{samel1}
\subsubsection{}
Let $\Gamma\subset G$ be a torsion-free geometrically finite discrete subgroup (see Definition \ref{t67}) such that all
its cusps are regular (Definition \ref{t799}). Furthermore let $({\varphi},V_{\varphi})$ be an admissible
twist (Definition \ref{weih134}). We are going to employ the notation introduced in Subsection \ref{geomf}.
\subsubsection{}
On $\bar Y$ we choose a partition of unity $\{\chi_p\}_{p\in {\mathcal{P}}_\Gamma\cup \{0\}}$ such that
$\chi_p\in C^\infty(\bar Y)$ and ${\mathrm{supp}}(\chi_p)\subset \bar{Y}_p$ for all $p$.
Restriction of these functions to $B_\Gamma$ gives a partition
of unity $\{\chi_p\}_{p\in{\mathcal{P}}_\Gamma^<\cup\{0\}}$ on $B_\Gamma$ such that
${\mathrm{supp}}(\chi_p)\subset B_p$. Here we denote the restriction
of $\chi_p$ to the boundary by the same symbol $\chi_p$.
In a similar manner for $P\in p\in {\mathcal{P}}^<_\Gamma$
we let $e_P:B_p\rightarrow
B_{U_P}$ denote the map defined as restriction of the map
$e_P:\bar Y_p\rightarrow \bar Y_{U_P}$.
Let $\chi_P\in C^\infty(\bar{Y}_{U_P})$ be the cut-off
function which is supported on
the range of $e_P$ and satisfies $\chi_p=e_P^*\chi_P$.
\subsubsection{}\label{neuj3005}
Using cut-off with $\chi_p$ and the map $e_P:B_p\rightarrow
B_{U_P}$ for $P\in p\in{\mathcal{P}}^<_\Gamma$ we define maps
\begin{eqnarray*}
T_P&:&C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))\rightarrow C^\infty(B_{U_P},V_{B_{U_P}}(\sigma_\lambda,{\varphi}))\\
T^P&:&C^\infty(B_{U_P},V_{B_{U_P}}(\sigma_\lambda,{\varphi}))\rightarrow C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))\ .
\end{eqnarray*}
\subsubsection{}
We define the Schwartz space for $\Gamma$ as the space of smooth sections
which belong to the Schwartz space \ref{weih245} near all cusps.
\begin{ddd}
We define ${S}_{\Gamma,k}(\sigma_\lambda,{\varphi})$ to be the subspace
of all $f\in C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$
such that $T_P(f)\in {S}_{U_P,k}(\sigma_\lambda,{\varphi})$ for all $P\in p\in {\mathcal{P}}_\Gamma^<$.
The inclusion $${S}_{\Gamma,k}(\sigma_\lambda,{\varphi})\hookrightarrow C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$
and the maps $$T_P:{S}_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow {S}_{U_P,k}(\sigma_\lambda,{\varphi})$$
induce on ${S}_{\Gamma,k}(\sigma_\lambda,{\varphi})$ the structure of a Fr\'echet space. Furthermore we define the Fr\'echet and Montel space
$${S}_{\Gamma}(\sigma_\lambda,{\varphi}):=\bigcap_{k\in{\mathbb {N}}_0}{S}_{\Gamma,k}(\sigma_\lambda,{\varphi})\ .$$
\end{ddd}
\subsubsection{}
\begin{lem}\label{loctrivg}
The families $\{{S}_{\Gamma,k}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$
and $\{{S}_{\Gamma}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$
form locally trivial holomorphic bundles.
\end{lem}
{\it Proof.$\:\:\:\:$}
Let $\Phi_{P,\lambda_0}$ denote the
trivialization constructed in Lemma \ref{ytra} which is
given by multiplication by $s_P^{(\lambda_0-\lambda)/\alpha}$
(here we add the index $P$ to the notation for the section $s$ constructed in Lemma \ref{ingf} in order to indicate its dependence on the parabolic subgroup).
Let $s_p:=T^P(s_P)$, where $P\in\tilde {\mathcal{P}}^<$ represents $p$.
Let $s_0\in C^\infty(B_\Gamma,V_{B_\Gamma}(1_{\rho+\alpha},{\varphi}))$
be any positive section (compare \ref{neuj911}).
Then we form $s_\Gamma:=\sum_{p\in {\mathcal{P}}_\Gamma} s_p + \chi_0 s_0$.
Multiplication by $s_\Gamma^{(\lambda_0-\lambda)/\alpha}$ defines isomorphisms
$$\Phi_{\Gamma,\lambda_0}: {S}_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow
{S}_{\Gamma,k}(\sigma_{\lambda_0},{\varphi})$$ and
$$\Phi_{\Gamma,\lambda_0}: {S}_{\Gamma}(\sigma_\lambda,{\varphi})\rightarrow
{S}_{\Gamma}(\sigma_{\lambda_0},{\varphi})\ .$$
We employ the maps $\Phi_{\Gamma,\lambda_0}$ in order to obtain the
required local trivializations and to implement the holomorphic structures.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Similarly to the case of the Schwartz space we define
the space $B_{\Gamma}(\sigma_\lambda,{\varphi})$
as the space of smooth sections with the same asymptotic expansions
near the cusps as the spaces as the space defined in \ref{neuj104}.
Recall the notation $\tilde {\mathcal{P}}^{max}$ from \ref{holle}.
\begin{ddd}
We define $$B_{\Gamma,k}(\sigma_\lambda,{\varphi}):=B_{\Gamma,k}(\sigma_\lambda,{\varphi})_1
\oplus \bigoplus_{P\in \tilde{\mathcal{P}}^{max}} {R}_{U_P}(\sigma_\lambda,{\varphi})\ ,$$
where $$B_{\Gamma,k}(\sigma_\lambda,{\varphi})_1\subset C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$ is
the subspace of all $f$ such that
$T_P(f)\in B_{U_P,k}(\sigma_\lambda,{\varphi})$ for all $P\in p\in {\mathcal{P}}_\Gamma^<$.
The map $$B_{\Gamma,k}(\sigma_\lambda,{\varphi})\hookrightarrow C^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))\ ,$$
the maps $$T_P:B_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow B_{U_P,k}(\sigma_\lambda,{\varphi})\ ,$$ $P\in \tilde {\mathcal{P}}^<$,
and the natural projections $$AS_P:B_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow {R}_{U_P}(\sigma_\lambda,{\varphi})\ ,$$ $P\in \tilde{\mathcal{P}}^{max}$,
equip $B_{\Gamma,k}(\sigma_\lambda,{\varphi})$ with the structure of a Fr\'echet
space.
We further define the Fr\'echet and Montel space
$$B_{\Gamma}(\sigma_\lambda,{\varphi}):=
\bigcap_{k\in{\mathbb {N}}_0} B_{\Gamma,k}(\sigma_\lambda,{\varphi})\ .$$
\end{ddd}
\subsubsection{}
We define $${R}_{\Gamma,k}(\sigma_\lambda,{\varphi}):=\bigoplus_{P\in \tilde{\mathcal{P}}} {R}_{U_P,k}(\sigma_\lambda,{\varphi}) \ ,\quad {R}_{\Gamma}(\sigma_\lambda,{\varphi}) :=\bigoplus_{P\in \tilde{\mathcal{P}}} {R}_{U_P}(\sigma_\lambda,{\varphi})\ .$$
We have asymptotic term maps $$AS_P: B_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow {R}_{U_P,k}(\sigma_\lambda,{\varphi})\ ,$$
$AS_P:=AS\circ T_P$, where $AS$ was defined in \ref{neuj915}
for pure cusps. The asymptotic term maps admit right-inverses
$$L_P:{R}_{U_P,k}(\sigma_\lambda,{\varphi})\rightarrow B_{\Gamma,k}(\sigma_\lambda,{\varphi})$$
given by the natural inclusion for $P\in\tilde{\mathcal{P}}^{max}$, and by $L_P:=T^P\circ L$
if $P\in \tilde{\mathcal{P}}^<$, where $L$ was defined in \ref{weih253}
for pure cusps.
Let $$AS:B_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow
{R}_{\Gamma,k}(\sigma_\lambda,{\varphi})$$ be induced by the maps $AS_P$ and $$L: {R}_{\Gamma,k}(\sigma_\lambda,{\varphi})
\rightarrow B_{\Gamma,k}(\sigma_\lambda,{\varphi})$$ be given by the sum of the maps
$L_P$ for the various $P\in\tilde{\mathcal{P}}$.
\subsubsection{}
\begin{lem}\label{spst}
The family $\{B_{\Gamma,k}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$
forms a trivial holomorphic bundle of Fr\'echet
spaces. We have a split (by $L$) exact sequence
\begin{equation}\label{trtra}0\rightarrow
{S}_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow B_{\Gamma,k}(\sigma_\lambda,{\varphi})
\stackrel{AS}{\rightarrow} {R}_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow 0\ .\end{equation}
Furthermore, the spaces $\{B_{\Gamma}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$ form a limit of locally trivial holomorpic bundles in the sense
of Subsection \ref{llim} and fit into the exact sequence (which does not admit any continuous split)
$$0\rightarrow {S}_{\Gamma}(\sigma_\lambda,{\varphi}) \rightarrow B_{\Gamma}(\sigma_\lambda,{\varphi}) \rightarrow
{R}_{\Gamma}(\sigma_\lambda,{\varphi})\rightarrow 0\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
The assertions follow from Lemma \ref{loctrivg} and the fact that the spaces ${R}_{\Gamma,k}(\sigma_\lambda,{\varphi})$ form locally trivial holomorphic vector bundles. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
We now show that the spaces $B_\Gamma(\sigma_\lambda,{\varphi})$
are compatible with twisting.
Let $(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$ be a finite-dimensional
representation of $G$ as in \ref{weih144}.
\begin{lem}\label{compat3}
\begin{enumerate}
\item
We have holomorphic families of continuous maps (see \ref{neuj1000} for notation)
$$i_{\sigma,\mu}^\Gamma:B_{\Gamma}(\sigma_\lambda,{\varphi})\rightarrow
B_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$$
and
$$p_{\sigma,\mu}^\Gamma:B_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow
B_{\Gamma}(\sigma_\lambda,{\varphi})\ .$$
\item
$B_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})$
is a ${\mathcal{Z}}({\mathfrak{g}})$-module.
\item
If $\lambda\not\in I_{\mathfrak{a}}$, then
$i_{\sigma,\mu}^\Gamma$ maps $B_{\Gamma}(\sigma_\lambda,{\varphi})$
isomorphically onto $$\ker_\Gamma(Z(\lambda)):=\ker\{Z(\lambda):
B_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow B_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})\}$$
(see \ref{weih146} for notation),
and the restriction of
$p_{\sigma,\mu}^\Gamma$ to
$$\ker\{\tilde Z(\lambda):B_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\rightarrow B_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})\}$$
is an isomorphism onto $B_{\Gamma}(\sigma_\lambda,{\varphi})$
(here $\tilde Z(\lambda)$ is the adjoint of
$\pi^{1_{-\lambda+\mu}\otimes \tilde\pi_{\sigma,\mu}, {\mathrm {id}}_{\tilde{\varphi}}}(\Pi(\lambda))$).
\item
The composition
$$j_{\sigma,\mu}^\Gamma:B_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes {\varphi})
\stackrel{1-Z(\lambda)}{\rightarrow} \ker_\Gamma(Z(\lambda))\stackrel{(i_{\sigma,\mu}^\Gamma)^{-1}}{\rightarrow}
B_{\Gamma}(\sigma_\lambda,{\varphi})$$
which is initially defined for $\lambda\not\in I_{\mathfrak{a}}$
extends to a meromorphic family of continuous maps.
Similarly, the composition
$$q_{\sigma,\mu}^\Gamma:B_{\Gamma}(\sigma_\lambda,{\varphi})
\stackrel{(p_{\sigma,\mu}^\Gamma)^{-1}}{\rightarrow} \ker_\Gamma(\tilde Z(\lambda)) \rightarrow B_\Gamma(1_{\lambda-\mu},\pi_{\sigma,\mu}\otimes {\varphi})$$
extends to a tame meromorphic family of continuous maps.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
In order to see 1. we employ Lemma \ref{compat2} 2., and
the identities $$i_{\sigma,\mu}^{U_P}\circ T_P=T_P\circ i_{\sigma,\mu}^\Gamma\ ,\quad
p_{\sigma,\mu}^{U_P}\circ T_P=T_P\circ p_{\sigma,\mu}^\Gamma\ .$$
\subsubsection{}
In order to see 2. we use Lemma \ref{compat2}, 3., and the fact
that $[\pi^{1_{\lambda+\mu}\otimes \pi_{\sigma,\mu},{\mathrm {id}}_{\varphi}}(A),\chi_p]$
is a differential operator with compactly supported coefficients on $B_\Gamma$
for any $A\in{\mathcal{Z}}({\mathfrak{g}})$.
\subsubsection{}
We show the first assertion of 3. and leave the second to the reader,
since the argument is similar.
Since $i^\Gamma_{\sigma,\mu}$ is ${\mathcal{Z}}({\mathfrak{g}})$-equivariant
it maps to $\ker_\Gamma(Z(\lambda))$ by construction of $Z(\lambda)$.
Let now $f\in\ker_\Gamma(Z(\lambda))$.
We claim that for any
$h\in C^\infty(B_\Gamma)$ we have $Z(\lambda)(hf)=0$.
Let $k$ be the order of the differential operator $Z(\lambda)$ which is a projection.
Then we have
\begin{eqnarray*}
Z(\lambda)(hf)&=& Z(\lambda)^{k+1}(hf)\\
&=&Z(\lambda)^{k}[Z(\lambda),h]f\\
&=&Z(\lambda)^{k-1}[Z(\lambda),[Z(\lambda),h]]f\\
&\dots&\\
&=&\underbrace{[Z(\lambda),\dots,[Z(\lambda),h]\dots]}_{k+1}f\\
&=&0\ .
\end{eqnarray*}
This shows that claim.
We conclude that $T_P(f)\in \ker_U(Z(\lambda))$ for
all $P\in\tilde P$.
By Lemma \ref{compat2} we find
$g_P\in B_U(\sigma_\lambda,{\varphi})$, $P\in\tilde{\mathcal{P}}$,
such that $T_P(f)=i_{\sigma,\mu}^{U_P}(g_P)$.
Let $$f_0:=f-\sum_{P\in\tilde {\mathcal{P}}} i_{\sigma,\mu}^\Gamma\circ T^P (g_P)\in
C_c^\infty(B_\Gamma,V_{B_\Gamma}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\ .$$
As in the proof of Lemma \ref{compat1},4. we can find
$g_0\in C_c^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$ such that
$f_0=i^\Gamma_{\sigma,\mu}(g_0)$.
Thus $f= i^\Gamma_{\sigma,\mu}\left(\sum_{P\in\tilde {\mathcal{P}}\cup\{0\}} g_P\right)$.
\subsubsection{}\label{neuj3006}
We show the first assertion of 4. and leave the second to the reader.
As in the proof of Lemma \ref{compat1}, 5. we construct
a holomorphic family
$$J:C_c^\infty(B_\Gamma,V_{B_\Gamma}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))
\rightarrow C_c^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$
as the composition $J:=j\circ (1-Z(\lambda))$,
where $$j:C_c^\infty(B_\Gamma,V_{B_\Gamma}(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))
\rightarrow C_c^\infty(B_\Gamma,V_{B_\Gamma}(\sigma_\lambda,{\varphi}))$$
comes from a bundle homomorphism.
Then we define
$$\tilde j^\Gamma_{\sigma,\mu}:= J \chi_0 (1-Z(\lambda))+
\sum_{P\in\tilde {\mathcal{P}}} \tilde T^P\circ j^{U_P}_{\sigma,\mu} \circ T_P\circ (1-Z(\lambda))\ .$$
Here $\tilde T^P$ is defined as $T^P$ but
using a cut-off function $\tilde\chi_p\in C^\infty(B_\Gamma)$, $P\in p$,
satisfying ${\mathrm{supp}}(\tilde\chi_p)\in B_p$ and $\tilde\chi_p\chi_p=\chi_p$
if $\tilde{\mathcal{P}}^<$. If $P\in\tilde{\mathcal{P}}^{max}$, then we set
$\tilde T^P:=T^P$. One checks that $\tilde j^\Gamma_{\sigma,\mu}\circ
i^\Gamma_{\sigma,\mu}={\mathrm {id}}$.
Since $\tilde j^\Gamma_{\sigma,\mu}$ vanishes on
$\ker(1-Z(\lambda))$ we conclude that it coincides
with $j^\Gamma_{\sigma,\mu}$ for $\lambda\not\in I_{\mathfrak{a}}$.
We thus have shown that $j^\Gamma_{\sigma,\mu}$
extends to a meromorphic family of continuous maps.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsection{Push-down}
\subsubsection{}
For the first part of the present subsection
we choose an Iwasawa decomposition
$G=KAN$ and a parabolic subgroup $P=MAN$, $M\subset K$.
Then we write $X=G/K$ and $\partial X=G/P=K/M$.
\subsubsection{}
For $f\in C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$
we consider the push-down
$$\pi^\Gamma_*(f)\in C^\infty(B_\Gamma,B_{B_\Gamma}(\sigma_\lambda,{\varphi}))\oplus \bigoplus_{P\in\tilde{\mathcal{P}}^{max}} {R}_{U_P}(\sigma_\lambda,{\varphi})\ .$$
It is given by the sum
$$(\pi^\Gamma_*(f))_1:=\sum_{g\in\Gamma}\pi^{\sigma_\lambda,{\varphi}}(g) (f_{|\Omega_\Gamma})\in{}^\Gamma C^\infty(\Omega_\Gamma,V(\sigma_\lambda,{\varphi}))$$
in the first component. Its second component
$$(\pi^\Gamma_*(f))_2\in \bigoplus_{P\in\tilde{\mathcal{P}}^{max}} {R}_{U_P}(\sigma_\lambda,{\varphi})$$
will be constructed below.
The goal of the present section is to show
that the push-down
converges for ${\rm Re }(\lambda)<-\delta_\Gamma-\delta_{\varphi}$ and defines
a holomorphic family of maps
$$\pi^\Gamma_*:C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow
B_{\Gamma}(\sigma_\lambda,{\varphi})\ .$$
\subsubsection{}
In the following lemma we consider points of $\partial X$ as subsets of $K$ using
the identification $\partial X\cong K/M$.
\begin{lem}\label{ggll}
If $W\subset\Omega_\Gamma$ is compact, then $\Gamma\cap WMA_+K$ is finite.
\end{lem}
{\it Proof.$\:\:\:\:$}
The set $ WMA_+K$ is a precompact subset of $X\cup \Omega_\Gamma$. Since
$\Gamma$ acts properly discontinuously on $X\cup \Omega_\Gamma$ the intersection of
the orbit $\Gamma K$ with $WMA_+K$ is finite. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
\begin{lem}\label{barpi}
For ${\rm Re }(\lambda)<-\delta_\Gamma-\delta_{\varphi}$
the sum $$\sum_{g\in\Gamma}\pi^{\sigma_\lambda,{\varphi}}(g)f_{|\Omega_\Gamma}$$ converges in $C^\infty(\Omega_\Gamma,V(\sigma_\lambda,{\varphi}))$ and defines a
holomorphic family of continuous maps
$$C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\to
{}^\Gamma C^\infty(\Omega_\Gamma,V(\sigma_\lambda,{\varphi}))\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
Using Lemma \ref{ggll} we employ the same argument as in the
proof of \cite{MR1749869}, Lemma 4.2. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Next we study the behaviour of the sum $\sum_{g\in\Gamma}\pi^{\sigma_\lambda,{\varphi}}(g)f_{|\Omega_\Gamma}$
near the cusps. Let $P\in\tilde{\mathcal{P}}$. We choose a system $\Gamma^P$ of representatives of $U_P\backslash \Gamma$
by taking in each class $[g]\in U_P\backslash \Gamma$ an element $h\in [g]$ which minimizes
${\rm dist}_X({\mathcal{O}},h{\mathcal{O}})$.
\begin{lem}\label{wwe}
There is a neighbourhood $W\subset\partial X$ of $\infty_P$ such that
$WMA_+K\cap \Gamma^P$ is finite.
\end{lem}
{\it Proof.$\:\:\:\:$}
Let $D({\mathcal{O}},U_P):=\{x\in X|{\rm dist}_X(x,{\mathcal{O}})\le {\rm dist}_X(gx,{\mathcal{O}})\:\forall g\in U_P\}$ be the
Dirichlet domain of $U_P$. If $h\in \Gamma^P$, then $h{\mathcal{O}}\in D({\mathcal{O}},U_P)$.
Let $\pi:D({\mathcal{O}},U_P)\rightarrow Y_{U_P}$ denote the projection and
consider $E:=\pi^{-1}(e_P(Y_p))$. Since
$e_P:Y_p\rightarrow Y_{U_P}$ is an isometry
we have
$\sharp(E\cap \Gamma^P{\mathcal{O}})\le 1$. Now the closure in $\bar X$ of
$D({\mathcal{O}},U_P)\setminus E$ does not contain $\infty_P$ and is therefore
disjoint from a neighbourhood $W\subset \partial X$ of $\infty_P$.
Since $WMA_+K\cap (D({\mathcal{O}},U_P)\setminus E)$ has a compact
closure inside $X$ we conclude that $WMA_+K\cap (D({\mathcal{O}},U_P)\setminus E)K\cap
\Gamma^P$ is finite. Since $\Gamma^P\subset D({\mathcal{O}},U_P)K$ we have shown the lemma.\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Let $P\in \tilde {\mathcal{P}}$ and $W_P$ be a neighbourhood of $\infty_P$ as constructed in Lemma \ref{wwe}.
Using Lemma \ref{wwe} and the arguments of the proof of
\cite{MR1749869}, Lemma 4.2, we show
\begin{lem}\label{bbarpi}
For ${\rm Re }(\lambda) < -\delta_\Gamma-\delta_{\varphi}$ and $f\in C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$ the sum
$$\bar\pi^P_*(f):=\sum_{g\in\Gamma^P}(\pi^{\sigma_\lambda,{\varphi}}(g)f)_{|W_P\cup\Omega_\Gamma}$$
converges in $C^\infty(W_P\cup\Omega_\Gamma,V(\sigma_\lambda,{\varphi}))$.
For varying $\lambda$ the maps $\bar\pi^P_*$ form a holomorphic family of continuous maps.
\end{lem} \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
We choose a cut-off function $\kappa_P\in C^\infty_c(W_P)$
such that $\kappa_P$ is equal to one near $\infty_P$.
Then $\kappa_P \bar\pi^P_*(f))\in C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$,
and we can write
$$(\pi^\Gamma_*(f))_1=\left((\pi^{U_P}_*(\kappa_P \bar\pi^P_*(f)))_1 + \sum_{u\in U_P} \pi^{\sigma_\lambda,{\varphi}}(u)(1-\kappa_P)
\bar\pi^P_*(f)\right)_{|\Omega_\Gamma}\ .$$
Here $$(\pi^{U_P}_*(\kappa_P \bar\pi^P_*(f)))_1=\pi^{U_P}_*(\kappa_P \bar\pi^P_*(f))$$
if the cusp associated to $U_P\subset P$ has smaller rank.
The second sum is locally finite and defines a smooth section on
$\Omega_\Gamma$. Moreover we have ($\bar \pi^P_*$ was defined in Lemma \ref{bbarpi})
$$\chi_P \sum_{u\in U_P} \pi^{\sigma_\lambda,{\varphi}}(u)(1-\kappa_P)
\bar\pi^P_*(f)_{|\Omega_\Gamma}\in C^\infty_c(B_{U_P},V_{B_{U_P}}(\sigma_\lambda,{\varphi}))\ .$$
\subsubsection{}
We define
$$(\pi^\Gamma_*(f))_2:=\bigoplus_{P\in\tilde{\mathcal{P}}^{max}} [\pi^{U_P}_*]\circ AS (\kappa_P \bar\pi^P_*(f))\ .$$
This definition is independent of the choice of $\kappa_P$.
Finally we define $\pi^\Gamma_*(f)$ to be the sum $(\pi^\Gamma_*(f))_1\oplus (\pi^\Gamma_*(f))_2$.
\subsubsection{}
\begin{lem}
For ${\rm Re }(\lambda) < -\delta_\Gamma-\delta_{\varphi}$ the push-down
$\pi^\Gamma_*$ induces a holomorphic family of continuous maps
$$\pi^\Gamma_*:C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))\rightarrow B_\Gamma(\sigma_\lambda,{\varphi})\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
This follows from Lemmas \ref{barpi}, \ref{bbarpi}, the observation that
\begin{equation}\label{xxxt}T_P(\pi^\Gamma_*(f))_1\in \chi_P \pi^{U_P}_*(\kappa_P \bar\pi^P_*(f)) +
C_c^\infty(B_{U_P},V_{B_{U_P}}(\sigma_\lambda,{\varphi}))\subset B_{U_P}(\sigma_\lambda,{\varphi})\end{equation}
for $P\in \tilde{\mathcal{P}}^<$, and the definition of $(\pi^\Gamma_*(f))_2$.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}\label{neuj1003}
Let $\chi^\Gamma\in C^\infty(\Omega_\Gamma)$ be a smooth cut-off function
such that $\sum_{g\in\Gamma} g^*\chi^\Gamma \equiv 1$.
We choose $\chi^\Gamma$ such that it coincides with $\chi^{U_P}$
in a neighbourhood of $\infty_P$, and such that
the restriction of the projection $\Omega_\Gamma\rightarrow B_\Gamma$
to ${\mathrm{supp}}(\chi^\Gamma)$ is proper.
Then multiplication by
$\chi^\Gamma$ defines a holomorphic family of right-inverses
of the push-down (see Lemma \ref{schwspl}) $$\{Q\}:{S}_\Gamma(\sigma_\lambda,{\varphi})\rightarrow
C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))$$
such that $\pi^{\Gamma}_*\circ \{Q\}={\mathrm {id}}$.
\subsubsection{}
Occasionally we will need the partial push-down
$$\pi^{\Gamma/U_P}_*:B_{U_P,k}(\sigma_\lambda,{\varphi})\rightarrow B_{\Gamma,k}(\sigma_\lambda,{\varphi})$$
which is defined as the average over $\Gamma/U_P$ if the cusp associated to
$U_P\subset P$ has smaller rank.
If the cusp associated to $U_P\subset P$ has full rank,
then we define
$$\pi^{\Gamma/U_P}_*(f\oplus v):=\pi^\Gamma_*(\{Q\}(f-(\pi^{U_P}_*\circ L\circ [Q](v))_1)+L \circ [Q](v))\ ,$$
where
$$f\oplus v\in C^\infty(B_{U_P},V_{B_{U_P}}(\sigma_\lambda,{\varphi}))\oplus {R}_{U_P}(\sigma_\lambda,{\varphi})=B_{U_P}(\sigma_\lambda,{\varphi})\ .$$
If $P\in \tilde{\mathcal{P}}^<$, then we have
$$\pi^{\Gamma/U_P}_*(f):=\pi^\Gamma_*(\{Q\}(f-\pi^{U_P}_*\circ L\circ [Q]\circ AS(f))+L \circ [Q] \circ AS (f))\ .$$
This formula together with (\ref{xxxt}) can be used to verify
that $\pi^{\Gamma/U_P}_*$ has the required mapping properties.
In particular we conclude that if
$P\in\tilde{\mathcal{P}}^<$, then the partial
push-down converges for
${\rm Re }(\lambda)<-\delta_\Gamma-\delta_{\varphi}$ and depends
holomorphically on $\lambda$. If $P\in\tilde{\mathcal{P}}^{max}$, then the partial push-down
is defined as a meromorphic family of continuous maps
for ${\rm Re }(\lambda)<-\delta_\Gamma-\delta_{\varphi}$.
\subsubsection{}
We collect the following useful identities
\begin{equation}\label{useid}\pi^{\Gamma/U_P}_*\circ \pi^{U_P}_*=\pi^\Gamma_*,\quad
\pi^{\Gamma/U_P}_* \circ T_P = \chi_p,\quad \pi^{\Gamma/U_P}_* \chi_P = T^P\ ,\end{equation}
where $P\in \tilde{\mathcal{P}}^<$ for the last two equations
and multiplication by $\chi_p$ implicitly involves
a projection onto the first component.
If $P\in \tilde{\mathcal{P}}^{max}$, then we have
\begin{equation}\label{useid2}
\pi^{\Gamma/U_P}_* \circ AS_P =AS_P
\ .\end{equation}
\subsubsection{}
We define $q_{\tilde\sigma,\mu}:=q^{\{1\}}_{\tilde\sigma,\mu}$ as in Lemma \ref{compat3}, 4. (applied to the trivial group).
\begin{lem}\label{compat5}
We have the following commutative
diagrams of meromorphic families of maps:
\begin{enumerate}
\item
$$\begin{array}{ccccc}
C^\infty(\partial X,V(\sigma_\lambda,{\varphi}))&\stackrel{i_{\sigma,\mu}}{\rightarrow}&C^\infty(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))&\stackrel{Z(\lambda)}{\rightarrow}&
C^\infty(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\\
\downarrow \pi^\Gamma_*&&\downarrow \pi^\Gamma_*&&\downarrow \pi^\Gamma_*\\
B_\Gamma(\sigma_\lambda,{\varphi})&\stackrel{i^\Gamma_{\sigma,\mu}}{\rightarrow}&
{\mathcal{B}}_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})&\stackrel{Z(\lambda)}{\rightarrow}&
{\mathcal{B}}_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})\end{array}\ .$$
\item
$$\begin{array}{ccc}
C^\infty(\partial X,V(\tilde\sigma_{-\lambda},\tilde{\varphi}))
&\stackrel{q_{\tilde\sigma,\mu}}{\rightarrow}&
C^\infty(\partial X,V(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi}))\\
\downarrow \pi^\Gamma_*&&\downarrow \pi^\Gamma_*\\
B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})
&\stackrel{p^\Gamma_{\tilde \sigma,\mu}}{\leftarrow}&
B_\Gamma(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi})\\
\end{array}\ .$$
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
The push-down is ${\mathcal{Z}}({\mathfrak{g}})$-equivariant.
This implies commutativity of the right square of the diagram 1.
The commutativity of the
left square is obvious for the part involving
$(\pi^\Gamma_*)_1$, and for $(\pi^\Gamma_*)_2$ we invoke
Definition \ref{cv1}.
For the second diagram consider the enlarged diagram
$$\xymatrix{C^\infty(\partial X,V(1_{-\lambda+\mu^\prime},\pi_{\tilde\sigma,\mu^\prime}\otimes\tilde{\varphi}))
\ar[d]^{\pi^\Gamma_*}&C^\infty(\partial X,V(\tilde\sigma_{-\lambda},\tilde{\varphi}))\ar@/_-0.5cm/[r]^{q_{\tilde\sigma,\mu}}
\ar[l]^{\qquad\quad i_{\tilde \sigma,\mu^\prime}}
\ar[d]^{\pi^\Gamma_*}&C^\infty(\partial X,V(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi}))
\ar[d]^{\pi^\Gamma_*}\ar[l]^{p_{\tilde \sigma,\mu}\quad}\\
B_\Gamma(1_{-\lambda+\mu^\prime},\pi_{\tilde\sigma,\mu^\prime}\otimes\tilde{\varphi})&
B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})
\ar[l]^{\qquad i^\Gamma_{\tilde \sigma,\mu^\prime}}&
B_\Gamma(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi})
\ar[l] ^{p^\Gamma_{\tilde \sigma,\mu}}}\ .$$
The left square commutes by the Definition \ref{cv1} of the push-down
in the middle. The outer rectangle commutes by the naturality of the definition of the push-down.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
Let $G^n$ be one of $\{Spin(1,n),SO(1,n)_0,SU(1,n),Sp(1,n)\}$.
Recall the definition of $i^*_\Gamma$ given in Subsection \ref{embedd}.
\begin{lem}\label{comppp}
We have the following commutative diagram:
$$\begin{array}{ccc}
C^\infty(\partial X^{n+m},V(1^{n+1}_{\lambda },{\varphi})) &\stackrel{i^*}{\rightarrow}&C^\infty(\partial X^n, V(1^n_{\lambda-\zeta},{\varphi})) \\
\downarrow \pi^{\Gamma,n+1}_*&&\downarrow \pi^{\Gamma,n}_* \\
B_\Gamma(1^{n+1}_{\lambda },{\varphi})&\stackrel{i_\Gamma^*}{\rightarrow }& B_\Gamma(1^n_{\lambda-\zeta},{\varphi})
\end{array}\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
We use Proposition \ref{klopp}.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsection{Extension, restriction, and the scattering matrix}
\subsubsection{}
We now define the distribution spaces $D_{\Gamma}(\sigma_\lambda,{\varphi})$ associated to $\Gamma$.
These spaces are the domain of the extension $ext^\Gamma$ and the Eisenstein series (see Definition \ref{neuj1001}).
\begin{ddd}
We define
$D_{\Gamma,k}(\sigma_\lambda,{\varphi})$ to be the dual space to
$B_{\Gamma,k}(\tilde\sigma_{-\lambda},\tilde{\varphi})$. Furthermore let
$D_{\Gamma}(\sigma_\lambda,{\varphi}):=\bigcup_{k\in{\mathbb {N}}_0} D_{\Gamma,k}(\sigma_\lambda,{\varphi})$.
\end{ddd}
Lemma \ref{spst} has the following consequence:
\begin{kor}
The spaces $\{D_{\Gamma,k}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$, form
a trivial holomorphic bundle of dual Fr\'echet spaces
and fit into the split exact sequence
$$0\rightarrow {R}_{\Gamma,k}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*\stackrel{AS^*}{\rightarrow} D_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow {S}_{\Gamma,k}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*\rightarrow 0\ .$$
The spaces $\{D_{\Gamma}(\sigma_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$,
are dual Fr\'echet and Montel spaces, and they
form a direct limit of trivial bundles. Furthermore
we have the exact sequence
$$0\rightarrow {R}_{\Gamma}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*\stackrel{AS^*}{\rightarrow} D_{\Gamma}(\sigma_\lambda,{\varphi})\rightarrow {S}_{\Gamma}(\tilde\sigma_{-\lambda},\tilde{\varphi})^*\rightarrow 0\ .$$
\end{kor}
Let $(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$ be a finite-dimensional
representation of $G$ as in Subsection \ref{ttw}.
\subsubsection{}
The following Lemma states that the distribution spaces are compatible with twisting.
\begin{lem}\label{compat6}
\begin{enumerate}
\item
$D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$
is a ${\mathcal{Z}}({\mathfrak{g}})$-module.
\item
There is a natural inclusion
$i^\Gamma_{\sigma,\mu}:D_\Gamma(\sigma_{\lambda},{\varphi})
\rightarrow D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$
which identifies
$D_\Gamma(\sigma_{\lambda},{\varphi})$ with $$\ker_\Gamma(Z(\lambda)):=\ker(Z(\lambda):
D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}) \rightarrow D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))$$
for $\lambda\not\in I_{\mathfrak{a}}$.
\item
Moreover, the map
$$j_{\sigma,\mu}^\Gamma:D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})
\stackrel{1-Z(\lambda)}{\longrightarrow} \ker_\Gamma(Z(\lambda))\stackrel{(i_{\sigma,\mu}^\Gamma)^{-1}}{\longrightarrow}
D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})$$
extends to a tame meromorphic family of continuous maps.
\end{enumerate}
\end{lem}
{\it Proof.$\:\:\:\:$}
We employ Lemma \ref{compat3}.
The map $i^\Gamma_{\sigma,\mu}$ is defined as the adjoint
of $$p^\Gamma_{\tilde\sigma,\mu}:B_\Gamma(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi})\rightarrow B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})\ .$$
$j_{\sigma,\mu}^\Gamma$ is just the adjoint
of $$q^\Gamma_{\tilde\sigma,\mu}:B_\Gamma(\tilde\sigma_{-\lambda},\tilde{\varphi})\rightarrow B_\Gamma(1_{-\lambda-\mu},\pi_{\tilde\sigma,\mu}\otimes\tilde{\varphi})\ .$$
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
\begin{ddd}For ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$
we define the extension map
$$ext^\Gamma:D_{\Gamma}(\sigma_\lambda,{\varphi})\rightarrow C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$$
to be the adjoint of the push-down
$$\pi^\Gamma_*:C^\infty(\partial X,V(\tilde\sigma_{-\lambda},\tilde{\varphi}))\rightarrow B_{\Gamma}(\tilde\sigma_{-\lambda},\tilde{\varphi})\ .$$
\end{ddd}
The extension
$ext^\Gamma$ is a holomorphic family of continuous maps and has values in $\Gamma$-invariant distributions.
We will also need the partial extensions
$ext^{\Gamma/U_P}:D_{\Gamma,k}(\sigma_\lambda,{\varphi})\rightarrow D_{U_P,k}(\sigma_\lambda,{\varphi})$,
which are defined as the adjoints of $\pi^{\Gamma/U_P}_*$.
Taking the adjoint of the relations (\ref{useid}) and (\ref{useid2})
we obtain
$$ext^{U_P}\circ ext^{\Gamma/U_P}=ext^\Gamma,\quad
T_P^*\circ ext^{\Gamma/U_P}=\chi_p,\quad \chi_P ext^{\Gamma/U_P} = (T^P)^*$$
(where $P\in\tilde{\mathcal{P}}^<$ for the last two, and multiplication by
$\chi_p$ implicitly involves projection onto the first component),
and
$$AS_P^*\circ ext^{\Gamma/U_P}=AS_P^*\ ,$$
if $P\in\tilde{\mathcal{P}}^{max}$.
\subsubsection{}
The space $D_{\{1\},k}(\sigma_\lambda,{\varphi})$ defined
in Def.~\ref{neuj1002} depends on the choice of a parabolic subgroup $P$.
In the present subsection we write $D_{P,k}(\sigma_\lambda,{\varphi})$
for that space and let $$D_{\{1\},k}(\sigma_\lambda,{\varphi}):=\bigcap_{p\in {\mathcal{P}}_\Gamma,P\in p} D_{P,k}(\sigma_\lambda,{\varphi})\ .$$
Using the relation $ext^{U_P}\circ ext^{\Gamma/U_P}=ext^\Gamma$
we see that $$ext^\Gamma: D_{\Gamma,k}(\sigma_\lambda,{\varphi})
\rightarrow D_{\{1\},k}(\sigma_\lambda,{\varphi})\ ,$$ and that it is holomorphic
as a map $ext^\Gamma: D_{\Gamma,k}(\sigma_\lambda,{\varphi})
\rightarrow D_{\{1\},k_1}(\sigma_\lambda,{\varphi})$ for
any $k_1>k$ (and, of course, for ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$).
\subsubsection{}\label{neuj3002}
From now on we consider the spherical case $\sigma=1$.
We introduce the restriction maps
$$res^\Gamma: D_{\{1\},k}(1_\lambda,{\varphi})\rightarrow D_{\Gamma,k_1}(1_\lambda,{\varphi})$$
defined for all $k$ and suitable $k_1\ge k$ depending on $k$.
The map $res^\Gamma$ is a refinement of the naive restriction
$$\{res^\Gamma\}: C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))\rightarrow
{S}_\Gamma(1_{-\lambda},\tilde{\varphi})^*$$ which is given
as the adjoint of $\{Q\}$ (see \ref{neuj1003}).
We define the meromorphic family of maps
$res^\Gamma$
by
\begin{equation}\label{rreessdef}
res^\Gamma(f):=\pi^\Gamma_*(\chi^\Gamma\chi_0 f) +
\sum_{P\in\tilde{\mathcal{P}}} T_P^* \circ res^{U_P}(f)\ ,\end{equation}
where we interprete $T_P^*$ as $AS_P^*$,
if $P\in\tilde{\mathcal{P}}^{max}$. Here $\pi^\Gamma_*$ simply stands for the average over $\Gamma$.
It is well-defined and holomorphic for all $\lambda\in{\aaaa_\C^\ast}$ since
$\chi^\Gamma\chi_0 f\in C_c^{-\infty}(\Omega_\Gamma,V(1_\lambda,{\varphi}))$
and thus $\pi^\Gamma_*(\chi^\Gamma\chi_0 f)\in C_c^{-\infty}(B_\Gamma,V_{B_\Gamma}(1_\lambda,{\varphi}))$.
Alternatively, one could write $\chi_0\{res^\Gamma\}(f)$ for $\pi^\Gamma_*(\chi^\Gamma\chi_0 f)$.
Like the maps $res^{U_P}$ (see Definition \ref{neuj707}) the restriction map $res^\Gamma$ depends on choices.
\subsubsection{}
A priori the composition $res^\Gamma\circ ext^\Gamma$
is meromorphic and depends on many choices. The following lemma shows that the situation is much better. It generalizes \ref{neuj2001}.
\begin{lem}\label{weih114}
The composition
$res^\Gamma\circ ext^\Gamma$
is regular and
coincides with the inclusion
of $$D_{\Gamma,k}(1_\lambda,{\varphi})\to D_{\Gamma,k_1}(1_\lambda,{\varphi})\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
Assume that $\lambda\in {\aaaa_\C^\ast}$ is such that $res^\Gamma$ is regular.
Note that we still assume ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$ in order ensure
convergence of $ext^\Gamma$.
Let $f\in D_{\Gamma,k}(1_\lambda,{\varphi})$.
It is easy to see that
$\pi^\Gamma_*(\chi^\Gamma\chi_0 ext^\Gamma(f))=\chi_0 f$.
Furthermore, if $P\in\tilde{\mathcal{P}}^<$, then
\begin{eqnarray*}
T_P^* \circ res^{U_P} \circ ext^\Gamma(f) &=&T_P^* \circ res^{U_P}
\circ ext^{U_P}\circ ext^{\Gamma/U_P}(f)\\
&=&T_P^* \circ ext^{\Gamma/U_P}(f)\\
&=&\chi_p f\ .
\end{eqnarray*}
If $P\in\tilde{\mathcal{P}}^{max}$, then
we have
\begin{eqnarray*}
AS_P^* \circ res^{U_P} \circ ext^\Gamma(f) &=&AS_P^* \circ res^{U_P}
\circ ext^{U_P}\circ ext^{\Gamma/U_P}(f)\\
&=&AS_P^* \circ ext^{\Gamma/U_P}(f)\\
&=&AS_P (f)\ .
\end{eqnarray*}
Summing these equations over $\tilde{\mathcal{P}}\cup\{0\}$, then we obtain
the desired identity $$res^\Gamma\circ ext^\Gamma(f)=f\ .$$
Since this equation holds true for generic $\lambda$, the assertion of the lemma follows. \hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
We also have the following useful identity
$$T_P^*\circ res^{U_P}= \chi_p res^\Gamma$$
for all $P\in p \in {\mathcal{P}}^<_\Gamma$.
\subsubsection{}\label{weih113}
Consider $P\in \tilde{\mathcal{P}}$ and
recall the Definition \ref{neuj108} of $Ext_{U_P}(1_\lambda,{\varphi})$.
If $\lambda$ is admissible (for $P$) in the sense of \ref{neuj300}, then on this space the restriction $res^{U_P}$ is well-defined independent of choices.
We now say that $\lambda\in {\aaaa_\C^\ast}$ is admissible, if it is admissible for all $P\in \tilde{\mathcal{P}}$.
Furthermore we define
$${}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))^0:=
{}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))\cap\bigcap_{P\in \tilde{\mathcal{P}}}
Ext_{U_P}(1_\lambda,{\varphi})\ .$$
It immediately follows from (\ref{rreessdef}) that $res^\Gamma$ is well-defined on this space independent of choices.
\subsubsection{}\label{neuj3001}
If we assume that $\lambda\not\in I_{\mathfrak{a}}$ or $\lambda>-\rho$, then
$${}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))^0={}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$$
by Prop.~\ref{maertins}, 2. Thus $res^\Gamma$ is pointwise well-defined at admissible $\lambda$
satisfying these additional conditions.
\subsubsection{}
\begin{ddd}
For ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$
we define the scattering matrix
$$S_\lambda^\Gamma:D_\Gamma(1_\lambda,{\varphi})\rightarrow D_\Gamma(1_{-\lambda},{\varphi})$$
as a meromorphic family of operators, which is given by
$$S_\lambda^\Gamma:=res^\Gamma \circ J_\lambda\circ ext^\Gamma$$
provided $-\lambda$ is non-integral and admissible in the sense of \ref{weih113}.
\end{ddd}
Using the scattering matrices for the cusps (Definition \ref{neuj2003}) we can rewrite the formula for $S^\Gamma_\lambda$ as follows:
\begin{equation}\label{spltt}
S_\lambda^\Gamma =\pi^\Gamma_* \circ\chi^\Gamma\chi_0 J_\lambda \circ ext^\Gamma + \sum_{P\in \tilde {\mathcal{P}}} T^*_P\circ S^{U_P}_\lambda \circ ext^{\Gamma/U_P}\ .\end{equation}
Using this formula and the results of
\ref{neuj2004}
we see that $S_\lambda^\Gamma$ is a well-defined
meromorphic family of continuous maps.
\subsubsection{}\label{gaga}
As in the case of pure cusps (Lemma \ref{fundd}) we obtain a natural inclusion
$$B_\Gamma(1_\lambda,{\varphi})\hookrightarrow D_\Gamma(1_\lambda,{\varphi})$$
provided that
$2\rho^{U_P}>l_{{\varphi}_{|U_P}}$ for all $P\in \tilde {\mathcal{P}}^<$.
Note that $l_{{\varphi}_{|U_P}}=2\delta_{{\varphi}_{|U_P}}$.
\begin{lem}\label{restrf}
We assume that all cusps of $\Gamma$ have smaller rank.
Assume that $\delta_{{\varphi}_{|U_P}}<\rho^{U_P}$ for all $P\in\tilde{\mathcal{P}}$.
Then the restriction of the scattering matrix to $B_\Gamma(1_\lambda,{\varphi})$
induces a meromorphic family of continuous maps
$$( S_\lambda^\Gamma)_{|B_\Gamma(1_\lambda,{\varphi})}:B_\Gamma(1_\lambda,{\varphi})\rightarrow B_\Gamma(1_{-\lambda},{\varphi})\ .$$
\end{lem}
{\it Proof.$\:\:\:\:$}
First note that if $f\in B_\Gamma(1_\lambda,{\varphi})$ (considered as distribution),
then $ext^\Gamma(f)$ is smooth on $\Omega_\Gamma$.
It follows that $J_\lambda\circ ext^\Gamma(f)$ is smooth on $\Omega_\Gamma$,
and hence
$S^\Gamma_\lambda(f)=res^\Gamma\circ J_\lambda\circ ext^\Gamma(f)$
is represented by a smooth function.
We choose a cut-off function $\tilde\chi_P$ on $B_{U_P}$ which is
supported in $e_{P}(B_p)$ such that $\tilde\chi_P\chi_P=\chi_P$.
In order to see that $S^\Gamma_\lambda(f)\in
B_\Gamma(1_{-\lambda},{\varphi})$ we employ
Equation (\ref{spltt}). Let $P\in\tilde{\mathcal{P}}$.
Then we have
\begin{eqnarray}
T_P\circ S^\Gamma_\lambda(f)&=&
T_P(\chi^\Gamma \chi_0 J_\lambda\circ ext^\Gamma(f))\\
&&+T_P\circ T_P^*\circ S^{U_P}_\lambda\circ ext^{\Gamma/U_P}(f)\nonumber\\
&=&T_P(\chi^\Gamma \chi_0 J_\lambda\circ ext^\Gamma(f))\label{ppo0}\\
&&+\chi_P^2 S_\lambda^{U_P}\circ \tilde\chi_P ext^{\Gamma/U_P}(f)\label{ppo1}\\
&&+\chi_P^2 S_\lambda^{U_P}\circ(1- \tilde\chi_P) ext^{\Gamma/U_P}(f)\label{ppo2}\ .
\end{eqnarray}
The term (\ref{ppo0}) belongs to $C_c^\infty(B_\Gamma,V_{B_\Gamma}(1_\lambda,{\varphi}))$.
We have $$\tilde\chi_P ext^{\Gamma/U_P}(f)=(\tilde T^P)^*(f)=\tilde T_P(f)\in
B_{U_P}(1_{-\lambda},{\varphi})\ ,$$ where $\tilde T_P, \tilde T^P$ are defined
as $T_P,T^P$, but using $\tilde\chi_P$ instead of $\chi_P$.
We can now apply Lemma \ref{inside} to (\ref{ppo1}) and Lemma \ref{offside} to (\ref{ppo2})
in order to see that these terms belong to $B_\Gamma(1_{-\lambda},{\varphi})$.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsection{Vanishing results}
\subsubsection{}
If $\lambda\in{\aaaa_\C^\ast}$ is admissible in the sense of \ref{weih113} and ${\rm Re }(\lambda)>-\rho$,
then $res^\Gamma$ is
well-defined on ${}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$
(see \ref{neuj3001}).
\subsubsection{}
Let $f\in {}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$.
The condition $\{res^\Gamma\}(f)=0$ (see \ref{neuj3002}) is equivalent to the condition that the support of $f$ as a distribution is contained in the limit set $\Lambda_\Gamma$.
In the case that $\Gamma$ is convex cocompact we know from \cite{MR1749869}, Thm.~4.7 that the space
$$\{f\in {}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))|\{res^\Gamma\}(f)=0\}$$
is trivial if ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$.
Already the example of pure cusps shows that this is not true in the presence of cusps.
Note that in the convex-cocompact case $\{res^\Gamma\}=res^\Gamma$.
It is at the heart of the matter that for geometrically finite groups $\Gamma$ and
${\rm Re }(\lambda)$ large the stronger condition
$res^\Gamma(f)=0$ implies $f=0$.
In \cite{MR1926489} we introduced a related condition ``$f$ is strongly supported on the limit set'' in order to show
such vanishing results.
\subsubsection{}
\begin{prop}\label{vani}
We assume that
$\lambda$ is admissible in the sense of \ref{weih113}
and satisfies
\begin{equation}\label{neuj3004}
{\rm Re }(\lambda)>\max\left(\{\delta_\Gamma\}\cup\{\rho_{U_P}\mid P\in\tilde{\mathcal{P}}\}\right)+\delta_{\varphi}
\ .
\end{equation}
Let $f\in {}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$. If
$res^\Gamma(f)=0$, then $f=0$.
\end{prop}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
Let $\lambda$ be admissible and $f\in {}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$. From the defining formula (\ref{rreessdef}) for $res^\Gamma$
we conclude that $res^\Gamma(f)=0$ if and only if $f$ is supported on $\Lambda_\Gamma$
and $res^{U_p}(f)\in C_c^{-\infty}(B_{U_P},V_{B_{U_P}}(1_\lambda,{\varphi}))\subset D_{U_P}(1_\lambda,{\varphi})$ for all $P\in\tilde{\mathcal{P}}$. The latter condition implies
that
\begin{equation}\label{wuff}
res^{U_P}(f)=\{res^{U_P}\}(f)
\end{equation}
for all $P\in\tilde{\mathcal{P}}$.
\subsubsection{}
If ${\rm Re }(\lambda)>-\rho^{U_P}$ and $h\in C^{\infty}(\partial X,V(1_{-\lambda},\tilde{\varphi}))$, then Proposition \ref{maertins}, 2, Lemma \ref{wauwau}, and Equation (\ref{wuff}) imply
$$
\langle f,h\rangle =\langle ext^{U_P}\circ res^{U_P}(f),h\rangle
=\langle \{res\}^{U_P}(f),\pi_*^{U_P}(h)\rangle $$
and therefore
\begin{equation}\label{woff}
\langle f,h\rangle =\sum_{u\in U_P} \langle\chi^{U_P}f,\tilde{\varphi}(u)h(u^{-1}.)\rangle\ .
\end{equation}
Thus, if $res^\Gamma(f)=0$, then $f$ is supported on the limit set and satisfies (\ref{woff})
for all $P\in\tilde{\mathcal{P}}$, $h\in C^{\infty}(\partial X,V(1_{-\lambda},\tilde{\varphi}))$. These are
precisely the defining conditions for being ``strongly supported on the limit set'' in the sense of \cite{MR1926489}. The main result of \cite{MR1926489} now states that an element
$f\in {}^\Gamma C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$ that is ``strongly supported on the limit set'' vanishes provided that (\ref{neuj3004}) holds. The proposition follows.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
\begin{kor}\label{wieder}
Assume $\lambda\in{\aaaa_\C^\ast}$ satisfies (\ref{neuj3004}). If
$f_\mu\in {}^\Gamma C^{-\infty}(\partial X,V(1_\mu,{\varphi}))$
is a germ of a meromorphic family at $\lambda$, then
$ext^\Gamma\circ res^\Gamma(f_\mu) = f_\mu$.
\end{kor}
{\it Proof.$\:\:\:\:$}
We apply $res^\Gamma$, and we use $res^\Gamma\circ ext^\Gamma={\mathrm {id}}$ and the injectivity
of $res^\Gamma$ for generic $\mu$ near $\lambda$ proved in Proposition \ref{vani}.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent.
\subsection{Meromorphic continuations and general $\sigma$}
\subsubsection{}
Recall that we have defined $ext^\Gamma$
for ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$.
The scattering matrix $S^\Gamma_\lambda$ was defined
on the same range of $\lambda$
for the trivial $M$-type
$1$.
In the present subsection we extend the definition
of the scattering matrix to general $M$-types.
Further we show that the extension
and the scattering matrix have meromorphic continuations
to all of ${\aaaa_\C^\ast}$.
This finishes the proof of theorems \ref{t119}, \ref{t110}, and \ref{t113}.
\subsubsection{}
We start with the meromorphic continuation of $S^\Gamma_\lambda$ to a certain half-plane
$W\subset {\aaaa_\C^\ast}$ in the spherical case. Under these conditions
we can construct a meromorphic family of parametrices
for $S_\lambda^\Gamma$ with finite-dimensional singularities.
In a second step we employ twisting in order to show
meromorphy of the extension and the scattering matrix on all of ${\aaaa_\C^\ast}$
and for general $\sigma$.
\subsubsection{}
\begin{prop}\label{meroext}
The extension
$$ext^\Gamma:D_\Gamma(1_\lambda,{\varphi})\rightarrow C^{-\infty}(\partial X,V(1_\lambda,{\varphi}))$$
and the scattering matrix $$S_\lambda^\Gamma:D_\Gamma(1_\lambda,{\varphi})\rightarrow
D_\Gamma(1_{-\lambda},{\varphi})$$ have meromorphic continuations to $$W:=\{\lambda\in{\aaaa_\C^\ast}\:|\:
{\rm Re }(\lambda)>-\rho+\beta\}\ ,$$ where $\beta=0$ for $X={H\mathbb{R}}^n, {H\mathbb{C}}^n$,
and $\beta=2\alpha$ for $X={H\mathbb{H}}^n$.
The family $ext^\Gamma$ has finite-dimensional singularities.
\end{prop}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
We first show the proposition under an additional assumption on $\Gamma$ and ${\varphi}$.
\begin{lem}\label{wahn1}
We assume that
\begin{equation}\label{kuckuck}
\delta_\Gamma+\delta_{\varphi}<-\max\left(\{0\}\cup\{\rho_{U_P}+\delta_{\varphi}\mid P\in\tilde{\mathcal{P}}\}\right)
\ .
\end{equation}
Then the extension $ext^\Gamma$
and the scattering matrix $S_\lambda^\Gamma$ have
a meromorphic continuation to the half-plane $W$.
The family $ext^\Gamma$ has finite-dimensional singularities.
\end{lem}
{\it Proof.$\:\:\:\:$}
\subsubsection{}
The additional assumption in particular implies that $\rho^{U_P}>\delta_{\varphi}\ge 0$
for all $P\in\tilde{\mathcal{P}}$.
Therefore
all cusps of $\Gamma$ have smaller rank, and
there is an embedding
$$B_\Gamma(1_\lambda,{\varphi})\hookrightarrow D_\Gamma(1_\lambda,{\varphi})$$
(see \ref{gaga}).
We consider the non-empty open subset
$$U:=\{\lambda\in{\aaaa_\C^\ast}\:|\:\max\{\delta_\Gamma+\delta_{\varphi},-\rho+\beta\}<{\rm Re }(\lambda)<\rho-\beta\}\subset {\aaaa_\C^\ast}\ .$$
Based on Lemma \ref{restrf} we first consider the restriction of the scattering matrix
to $B_\Gamma(1_\lambda,{\varphi})$.
For $\lambda\in U$ we construct a parametrix (see \ref{neuj3005} for $T_P$ and \ref{neuj3006} for $\tilde T^P$)
$$Q_{-\lambda}:B_\Gamma(1_{-\lambda},{\varphi})\rightarrow B_\Gamma(1_\lambda,{\varphi})$$
of $S^\Gamma_\lambda$ by
$$Q_{-\lambda} (f) := \pi^\Gamma_*(\chi^\Gamma \chi_0 J_{-\lambda} (\chi^\Gamma
\tilde\chi_0 f)) + \sum_{P\in\tilde{\mathcal{P}}} T_P^* \circ S_{-\lambda}^{U_P}\circ (\tilde T^P)^* (f)\ .$$
Here $\tilde \chi_0$ is some cut-off function on $B_\Gamma$
of compact support such that $\chi_0\tilde \chi_0=\chi_0$, and
$\tilde T^P$ is defined in the same way as
$T^P$ but using a cut-off function $\tilde\chi_p$ with support
on $B_p$ such that $\chi_p\tilde\chi_p=\chi_p$.
Note that $Q_{-\lambda}$ has a continuous extension
to a map
$$Q_{-\lambda}:D_\Gamma(1_{-\lambda},{\varphi})\rightarrow D_\Gamma(1_\lambda,{\varphi})\ .$$
This is clear for the second term $\sum_{P\in\tilde{\mathcal{P}}} T_P^* \circ S_{-\lambda}^{U_P}\circ (\tilde T^P)^*$
since the scattering matrices $S_{-\lambda}^{U_P}$ extend to distributions.
The map $f\mapsto\chi^\Gamma\chi_0J_{-\lambda}(\chi^\Gamma\tilde\chi_0f)$
extends to a continuous map from $D_\Gamma(1_{-\lambda},{\varphi})$ to
$C^{-\infty}_c(\Omega_\Gamma,V(1_\lambda,{\varphi}))$.
The push-down extends to a continuous map from
$C^{-\infty}_c(\Omega_\Gamma,V(1_\lambda,{\varphi}))$
to $C^{-\infty}_c(B_\Gamma,V_{B_\Gamma}(1_\lambda,{\varphi}))$.
This implies that the first term
$\pi^\Gamma_*(\chi^\Gamma \chi_0 J_{-\lambda} (\chi^\Gamma \tilde\chi_0 f))$
extends to distributions, too.
\subsubsection{}
We claim that for $|{\rm Re }(\lambda)|<\rho-\beta$ the scattering matrix $S_\lambda^{U_P}$
has at most finite-dimensional singularities. To see the claim we write
$S_\lambda^{U_P}=res^{U_P}\circ J_\lambda\circ ext^{U_P}$.
Since $J_\lambda$ is regular and bijective for these
$\lambda$, and $ext^{U_P}$ has finite-dimensional singularities,
the only term which may contribute infinite-dimensional singularities
is $res^{U_P}$.
Since $\{res^{U_P}\}$ is always regular the singular part of $res^{U_P}$ has values
in the space ${R}_{U_P}(1_{\lambda},\tilde{\varphi})^*$.
Now $S_\lambda^{U_P}$ is the continuous extension of its restriction
to the dense subspace $B_{U_P}(1_\lambda,{\varphi})$.
Since $S^{U_P}_\lambda$ maps $B_{U_P}(1_\lambda,{\varphi})$ to $B_{U_P}(1_{-\lambda},{\varphi})$
the range of the singular part of $S_\lambda^{U_P}$ does not
contain non-trivial elements of ${R}_{U_P}(1_{\lambda},\tilde{\varphi})^*$.
Let $D_{U_P}(1_\lambda,{\varphi})^0\subset D_{U_P}(1_\lambda,{\varphi})$
be a closed subspace of finite codimension on which $ext^{U_P}$
is regular and put $B_{U_P}(1_\lambda,{\varphi})^0:=
D_{U_P}(1_\lambda,{\varphi})^0\cap B_{U_P}(1_\lambda,{\varphi})$.
Then $B_{U_P}(1_\lambda,{\varphi})^0$ has finite codimension
in $B_{U_P}(1_\lambda,{\varphi})$, and the closure of
$B_{U_P}(1_\lambda,{\varphi})^0$ in $D_{U_P}(1_\lambda,{\varphi})$ is
$D_{U_P}(1_\lambda,{\varphi})^0$.
Looking at Laurent expansions we conclude that the restriction of $S_\lambda^{U_P}$ to $B_{U_P}(1_\lambda,{\varphi})^0$ and hence to
$D_{U_P}(1_\lambda,{\varphi})^0$ is regular.
This proves the claim.
We conclude that for
$\pm\lambda\in U$ the family $Q_\lambda$ has at most finite-dimensional
singularities.
\subsubsection{}
We define the remainder
$$R_\lambda:=Q_{-\lambda}\circ S^\Gamma_\lambda-{\mathrm {id}} \ .$$
We are going to show that $R_\lambda$ is a meromorphic family of
smoothing operators with finite-dimensional singularities.
We start with
\begin{eqnarray*}
R_\lambda(f)&=&Q_{-\lambda}\circ S^\Gamma_\lambda(f)-f\\
&=&\pi^\Gamma_* (\chi^\Gamma \chi_0 J_{-\lambda}( \chi^\Gamma\tilde\chi_0 S^\Gamma_\lambda(f)))\\
&& + \sum_{P\in\tilde{\mathcal{P}}} T_P^* S_{-\lambda}^{U_P}(\tilde T^P)^* S^\Gamma_\lambda(f)-f
\end{eqnarray*}
Next we insert the definition $S^\Gamma_\lambda=res^\Gamma\circ J_\lambda\circ
ext^\Gamma$ and the definition of $res^\Gamma$ (\ref{rreessdef}). We obtain
\begin{eqnarray*}
R_\lambda(f)&=&
\pi^\Gamma_*(\chi^\Gamma \chi_0 J_{-\lambda} (\chi^\Gamma\tilde\chi_0 \pi^\Gamma_*(\chi^\Gamma\chi_0 J_\lambda\circ ext^{\Gamma}(f)))) \\
&&+\sum_{P\in\tilde{\mathcal{P}}}\pi^\Gamma_* (\chi^\Gamma \chi_0 J_{-\lambda} (\chi^\Gamma\tilde\chi_0 T_P^* res^{U_P}\circ J_\lambda \circ ext^{\Gamma}(f)))\\
&&+\sum_{P\in\tilde{\mathcal{P}}} T_P^* S_{-\lambda}^{U_P}
(\tilde T^P)^*\pi^\Gamma_*( \chi^\Gamma\chi_0 J_\lambda \circ ext^\Gamma(f))\\
&&+\sum_{P,Q\in\tilde{\mathcal{P}}} T_P^* S_{-\lambda}^{U_P} (\tilde T^P)^*
T_Q^* (res^{U_Q}\circ J_\lambda \circ ext^{\Gamma} (f)) - f\ .
\end{eqnarray*}
Now we employ
\begin{eqnarray*}
\chi^\Gamma\tilde\chi_0 \pi^\Gamma_* \chi^\Gamma \chi_0&=&
\chi^\Gamma\chi_0\\
\chi^\Gamma\tilde\chi_0\sum_{P\in\tilde{\mathcal{P}}} T_P^*\circ res^{U_P}
&=& \tilde\chi_0 (1-\chi_0)\chi^\Gamma\\
\sum_{Q\in\tilde{\mathcal{P}}} (\tilde T^P)^*
T_Q^* res^{U_Q} + (\tilde T^P)^*\pi^\Gamma_* \chi^\Gamma\chi_0 & =& \tilde \chi_P res^{U_P}\\
res^{U_P}\circ J_\lambda\circ ext^\Gamma &=& S^{U_P}_\lambda\circ ext^{\Gamma/U_P}
\end{eqnarray*} in order
to obtain
\begin{eqnarray*}
R_\lambda(f)
&=&\pi^\Gamma_*(\chi^\Gamma \chi_0 J_{-\lambda} (\chi^\Gamma\tilde\chi_0 J_\lambda \circ ext^{\Gamma}(f))) \\
&&+\sum_{P\in\tilde{\mathcal{P}}} T_P^* S_{-\lambda}^{U_P} (\tilde\chi_P
S_{\lambda}^{U_P} \circ ext^{\Gamma/U_P}(f))- f\ .
\end{eqnarray*}
Using the functional equations of the intertwining operators and scattering matrices we can further write
\begin{eqnarray*}
R_\lambda(f)
&=&\pi^\Gamma_*(\chi^\Gamma \chi_0 ext^{\Gamma}(f)) \\
&&-\pi^\Gamma_*(\chi^\Gamma \chi_0 J_{-\lambda}( (1-\chi^\Gamma\tilde\chi_0) J_\lambda \circ ext^{\Gamma}(f))) \\
&&+\sum_{P\in\tilde{\mathcal{P}}} T_P^*
ext^{\Gamma/U_P}(f)\\
&&-\sum_{P\in\tilde{\mathcal{P}}} T_P^* S_{-\lambda}^{U_P} ((1-\tilde\chi_P)
S_{\lambda}^{U_P} \circ ext^{\Gamma/U_P}(f))- f\\
&=&-\pi^\Gamma_*(\chi^\Gamma \chi_0 J_{-\lambda}( (1-\chi^\Gamma\tilde\chi_0) J_\lambda\circ ext^{\Gamma}(f))) \\
&&-\sum_{P\in\tilde{\mathcal{P}}} T_P^* S_{-\lambda}^{U_P} ((1-\tilde\chi_P)
S_{\lambda}^{U_P}\circ ext^{\Gamma/U_P}(f))\ .
\end{eqnarray*}
Note that $\chi^\Gamma \chi_0 (1-\chi^\Gamma\tilde\chi_0)=0$,
and that $J_{-\lambda}$ is off-diagonal smoothing. Furthermore note
that $\chi_P (1-\tilde\chi_P)=0$, and that $S_{-\lambda}^{U_P}$
is off-diagonal smoothing by Lemma \ref{offside}.
We conclude that
$R_\lambda$ extends to a meromorphic family of continuous maps
$$R_\lambda:D_\Gamma(1_\lambda,{\varphi})\rightarrow B_\Gamma(1_\lambda,{\varphi})$$
with finite-dimensional singularities.
Since $J_0={\mathrm {id}}$ and $S_0^{U_P}={\mathrm {id}}$ we have $R_0=0$.
\subsubsection{}
We define the Banach space ${S}_{\Gamma,0,0}(1_\lambda,{\varphi})$ to be the subspace
of all $f\in C(B_\Gamma,V_{B_\Gamma}(1_\lambda,{\varphi}))$
such that $T_P(f)\in {S}_{U_P,0,0}(1_\lambda,{\varphi})$ for all $P\in \tilde{\mathcal{P}}$ (see \ref{weih259}).
Then we have compact inclusions
$$ B_\Gamma(1_\lambda,{\varphi})\hookrightarrow {S}_{\Gamma,0,0}(1_\lambda,{\varphi})
\hookrightarrow D_\Gamma(1_\lambda,{\varphi})$$
(in order to see the second inclusion use Lemma \ref{fundd}).
The family of Banach spaces $\{{S}_{\Gamma,0,0}(1_\lambda,{\varphi})\}_{\lambda\in{\aaaa_\C^\ast}}$
forms a trivial holomorphic bundle.
\subsubsection{}
We now apply meromorphic Fredholm theory \cite{reedsimon78}, Thm. VI.13,
(or better its version for families of operators on Banach spaces)
to the family $$1+R_\lambda:{S}_{\Gamma,0,0}(1_\lambda,{\varphi})\rightarrow
{S}_{\Gamma,0,0}(1_\lambda,{\varphi})\ .$$
As in \cite{MR1749869}, Lemma 3.6, we conclude that $(1+R_\lambda)^{-1}$ exists as a meromorphic
family of maps on ${S}_{\Gamma,0,0}(1_\lambda,{\varphi})$ of the form
$(1+T_\lambda)$, where $T_\lambda$ is a meromorphic family of continuous
operators from $D_\Gamma(1_\lambda,{\varphi})$ to $B_\Gamma(1_\lambda,{\varphi})$
with finite-dimensional singularities.
\subsubsection{}
We see that the inverse of the scattering matrix is given on $U$ by $$(S^\Gamma_\lambda)^{-1} := (1+R_\lambda)^{-1} Q_{-\lambda}\ .$$
as a meromorphic family and with finite-dimensional singularities.
\subsubsection{}
We have $-U\cup \{\lambda\in{\aaaa_\C^\ast}\:|\: {\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}\}=W$.
By our assumption the set
$$W_0=\left\{\lambda\in{\aaaa_\C^\ast}\:|\: \delta_\Gamma+\delta_{\varphi} <
{\rm Re }(\lambda)< -\delta_{\varphi}-\max\left(\{\delta_\Gamma\}\cup\{\rho_{U_P}\mid P\in\tilde{\mathcal{P}}\}\right)\right\}$$
is non-empty.
By Corollary \ref{wieder} we find for $\lambda\in W_0$ the meromorphic identity
\begin{eqnarray*}
S_{-\lambda}^\Gamma\circ S^\Gamma_\lambda&=&
res^\Gamma\circ J_{-\lambda}\circ ext^\Gamma\circ res^\Gamma\circ J_\lambda\circ ext^\Gamma\\
&=&res^\Gamma\circ J_{-\lambda}\circ J_\lambda\circ ext^\Gamma\\
&=&res^\Gamma\circ ext^\Gamma\\
&=&{\mathrm {id}} \ .
\end{eqnarray*}
Thus defining $S^\Gamma_\lambda:=(S_{-\lambda}^\Gamma)^{-1}$ for $-\lambda\in U$
we obtain the continuation of $S^\Gamma_\lambda$ to $W$.
\subsubsection{}
We obtain the meromorphic continuation of $ext^\Gamma$ to $W$
with finite-dimensional singularities using the identity
$$ext^\Gamma=J_{-\lambda}\circ ext^\Gamma\circ S_\lambda^\Gamma\ .$$
See the proof of \cite{MR1749869}, Lemma 5.11,
for the corresponding argument.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
The following lemma finishes the proof of Proposition \ref{meroext}.
\begin{lem}\label{hjdhdjwqdwqdwqdwq}
Lemma \ref{wahn1} holds true without the assumption (\ref{kuckuck}).
\end{lem}
{\it Proof.$\:\:\:\:$}
We are going to employ the embedding trick.
First we assume that $G^n$ belongs to the list
$\{Spin(1,n),SO(1,n)_0,SU(1,n),Sp(1,n)\}$.
If we consider $\Gamma\subset G^n$ as a subgroup of $G^{n+m}$,
then for sufficiently large $m$ the inequality (\ref{kuckuck})
is satisfied. Indeed, all ingredients of (\ref{kuckuck}) but $\delta_\Gamma$
are independent of $m$, while $\delta_\Gamma^{n+m}=\delta^n_\Gamma-m\zeta$, $\zeta=\rho^{n+1}-\rho^n$.
It thus suffices to show that a meromorphic continuation of $ext^{\Gamma,n+m}$
implies a meromorphic continuation of $ext^{\Gamma,n}$.
Assume that
we have a meromorphic continuation of $ext^{\Gamma,n+m}$
to $W^{n+m}:=\{\lambda\in{\aaaa_\C^\ast}\:|\:
{\rm Re }(\lambda)>-\rho^{n+m}+\beta\}$ with finite-dimensional singularities.
Then we employ the diagram
$$\begin{array}{ccc} C^{-\infty}(\partial X,V(1^{n}_\lambda,{\varphi})) &
\stackrel{ j^*}{ \leftarrow} &C^\infty(\partial X^{n+m}, V(1^{n+m}_{\lambda-m\zeta},{\varphi}))
\\
\uparrow ext^{\Gamma,n}_*&&\uparrow ext^{\Gamma,n+m}_* \\ D_\Gamma(1^{n}_\lambda,{\varphi})&\stackrel{i^\Gamma_*}{\rightarrow }&D_\Gamma(1^{n+m}_{\lambda-m\zeta},{\varphi})
\end{array}$$
which apriori holds in the domain of convergence
${\rm Re }(\lambda)>\delta_\Gamma^n+\delta_{\varphi}^n$ by Lemma \ref{comppp}. Note that the left-inverse $j^*$ of $i^*$ was constructed in Lemma \ref{jjjjjjjjjjjtwzew} (or better its iterate for the $m$-fold embedding).
We can define the meromorphic continuation to
$W^n=W^{n+m}+m\zeta$
of $ext^{\Gamma,n}$ using the identity $ext^{\Gamma,n}=j^*\circ ext^{\Gamma,n+m}\circ i^\Gamma_*$.
Let now $G$ be general. Then we can find a subgroup
$\Gamma^0\subset \Gamma$ of finite index to which we can apply
the concept of embedding (see \ref{ohu}). In particular we obtain a
meromorphic continuation of
$ext^{\Gamma^0}$ to $W$. If we identify
$D_\Gamma(1_\lambda,{\varphi})$ with the subspace
${}^{\Gamma/\Gamma^0} D_{\Gamma^0}(1_\lambda,{\varphi})$,
then we obtain the meromorphic continuation
of $ext^\Gamma$ using the identity
$ext^{\Gamma}=ext^{\Gamma^0}_{|D_\Gamma(1_\lambda,{\varphi})}$.
Using the meromorphic continuation of $ext^{\Gamma,n}$
we then obtain the meromorphic continuation of
the scattering matrix $S^\Gamma_\lambda$ by
$S_\lambda^\Gamma=res^\Gamma\circ J_\lambda\circ ext^\Gamma$.
\hspace*{\fill}$\Box$ \\[0.5cm]\noindent
\subsubsection{}
In Proposition \ref{meroext} we have obtained (in the spherical case) a meromorphic extension of
$ext^\Gamma$ and the scattering matrix $S_\lambda^\Gamma$
to a half space $W\subset {\aaaa_\C^\ast}$. In the following argument, using twisting, we find the meromorphic continuation of these objects to all of ${\aaaa_\C^\ast}$.
\subsubsection{}
Let $\sigma$ be a Weyl invariant representation of $M$,
and $(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$ be a finite-dimensional
representation of $G$ as in
\ref{weih144}.
The adjoint of the diagram
2. of Lemma \ref{compat5}
gives
the following commutative diagram of
meromorphic families of maps
\begin{equation}\label{dieletzte}\begin{array}{ccc}
D_\Gamma(\sigma_\lambda,{\varphi})
&\stackrel{i^\Gamma_{\sigma,\mu}}{\rightarrow}&D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})\\
\downarrow ext^\Gamma&&\downarrow ext^\Gamma\\
C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))
&\stackrel{j_{\sigma,\mu}}{\leftarrow}&
C^{-\infty}(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}
\otimes{\varphi}))
\end{array}
\end{equation}
for ${\rm Re }(\lambda)>\delta_\Gamma+\delta_{\varphi}$.
Using Proposition \ref{meroext} we conclude that
$$ext^\Gamma:D_\Gamma(\sigma_\lambda,{\varphi})\rightarrow C^{-\infty}(\partial X,V(\sigma_\lambda,{\varphi}))$$
is meromorphic on the set $W-\mu$.
Since we can choose $\mu$ arbitrary large
we obtain the meromorphic continutaion of $ext^\Gamma$ to all of ${\aaaa_\C^\ast}$.
From (\ref{dieletzte}),
injectivity of $i_{\sigma,\mu}$,
and Proposition \ref{meroext} we conclude that $ext^\Gamma$
has finite-dimensional singularities. This finishes the proof of Theorem \ref{t110}.
The adjoint of the extension is the push-down.
Therefore Theorem \ref{t119} follows from \ref{t110}.
\subsubsection{}
Let for a moment be $\sigma=1$. Using the meromorphic continuation of
the extension we obtain the meromorphic continuation of the
scattering matrix to all of ${\aaaa_\C^\ast}$
by the formula $S^\Gamma_\lambda=res^\Gamma\circ J_\lambda\circ ext^\Gamma$.
In order to proceed similarly for general $\sigma$ we first have to define the
restriction map in this situation.
Let now $\sigma$ be a Weyl-invariant representation of
$M$
and $(\pi_{\sigma,\mu},V_{\pi_{\sigma,\mu}})$ be a finite-dimensional
representation of $G$ as in
\ref{weih144}.
For $k\in {\mathbb {N}}$
we define restriction maps
$$res^\Gamma:{}^\Gamma C^{-k}(\partial X,V(\sigma_\lambda,{\varphi}))
\rightarrow D_\Gamma(\sigma_\lambda,{\varphi})$$
using the diagrams
$$\begin{array}{ccc}
C^{-k}(\partial X,V(\sigma_\lambda,{\varphi}))
&\stackrel{i_{\sigma,\mu}}{\rightarrow}&C^{-k}(\partial X,V(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi}))\\
\downarrow res^\Gamma&&\downarrow res^\Gamma\\
D_\Gamma(\sigma_\lambda,{\varphi})&\stackrel{j^\Gamma_{\sigma,\mu}}{\leftarrow}&
D_\Gamma(1_{\lambda+\mu},\pi_{\sigma,\mu}\otimes{\varphi})
\end{array}\ .$$
The right column depends on choices (including $\mu$), and so does
the left column. But it is easy to check that
\begin{equation}\label{juhu}
res^\Gamma\circ ext^\Gamma ={\mathrm {id}}
\end{equation}
\subsubsection{}\label{hjshqwjsdwhqdwqdwqdwqdqw}
Note that (\ref{dieletzte}) implies the meromorphic identity $ext^\Gamma\circ j_{\sigma,\mu}^\Gamma=j_{\sigma,\mu}\circ ext^\Gamma$.
Let $f_\nu\in {}^\Gamma C^{-\infty}(\partial X,V(\sigma_\nu,{\varphi}))$
be a meromorphic family defined on {\em all} of ${\aaaa_\C^\ast}$.
We compute for $\nu\gg 0$ using Corollary \ref{wieder}
and any set of choices for $res^\Gamma$
\begin{eqnarray}\label{jetztaber}
ext^\Gamma\circ res^\Gamma(f_\nu)&=&
ext^\Gamma\circ j_{\sigma,\mu}^\Gamma \circ res^\Gamma\circ i_{\sigma,\mu}(f_\nu)\nonumber\\
&=&j_{\sigma,\mu}\circ ext^\Gamma\circ res^\Gamma\circ i_{\sigma,\mu}(f_\nu)\nonumber\\
&=&j_{\sigma,\mu}\circ i_{\sigma,\mu}(f_\nu)\nonumber\\
&=&f_\nu\ .
\end{eqnarray}
Formula (\ref{juhu}) implies that $ext^\Gamma$ is injective
on families.
Thus we can conclude from (\ref{jetztaber}) that $res^\Gamma(f_\nu)$ is independent of choices.
\subsubsection{}
Now we can define the scattering matrix for general
$\sigma$ by the same formula as in the spherical case
$$S_\lambda^\Gamma:=res^\Gamma\circ J_\lambda\circ ext^\Gamma\ .$$
It is meromorphic on ${\aaaa_\C^\ast}$.
In order to finish the proof of Theorem \ref{t113}
it remains to show the functional equation.
Using (\ref{jetztaber}) we find
\begin{eqnarray*}
S^\Gamma_{-\lambda}\circ S_\lambda^\Gamma&=&
res^\Gamma\circ J_{-\lambda} \circ ext^\Gamma\circ
res^\Gamma\circ J_\lambda\circ ext^\Gamma\\
&=&res^\Gamma\circ J_{-\lambda}\circ J_\lambda\circ ext^\Gamma\\
&=&res^\Gamma\circ ext^\Gamma\\
&=&{\mathrm {id}}\ .
\end{eqnarray*}
\bibliographystyle{plain}
|
\section{Introduction}
It has long been believed that trans-Alfv\'{e}nic shock waves (TASWs), at
which the flow velocity passes over the Alfv\'{e}n velocity, cannot exist in
the real world. Since a stationary trans-Alfv\'{e}nic shock transition was
obtained in a numerical simulation \cite{wu87}, this conventional view point
was replaced by an opposite view point. The overall claim was that there is
no principal difference between TASWs and fast and slow shocks, at which the
flow is super- and sub-Alfv\'{e}nic, respectively. At the same time, the
contradiction inherent in a stationary TASW, which follows from an
analytical theory, was not lifted. To reconcile this contradiction, it was
suggested that a TASW exists in an unsteady state in which it is repeatedly
destroyed and recovered \cite{m98}. In the present paper, we show by way of
magnetohydrodynamic (MHD) simulation that the evolution of a TASW may have
the form of oscillatory disintegration, i.e., reversible transformation into
another TASW.
The disintegration of an arbitrary hydrodynamic discontinuity was considered
for the first time by Kotchine \cite{kotchine}. After that, Bethe
\cite{bethe} studied the disintegration of shock waves. In the absence of a
magnetic field, the shock may disintegrate only in a medium with anomalous
thermodynamic properties. The magnetic field enlarges the number of possible
discontinuous structures thus giving additional degrees of freedom for the
disintegration. The disintegration configurations of arbitrary MHD
discontinuities were obtained in Refs. \cite{arbit}. Furthermore, it has
been shown that trans-Alfv\'{e}nic shock transitions can be realized also
through a set of several discontinuities \cite{shock}, in contrast with fast
and slow transitions. However, this fact on its own does not assure that the
shock disintegrates.
The important feature that predetermines the disintegration of TASWs is their
nonevolutionarity. The problem of evolutionarity was initially formulated for
the fronts of combustion \cite{landau} and hydrodynamic
discontinuities \cite{cf}. Evolutionarity is a
property of a discontinuous flow to evolve in such a way that the flow
variation remains small under the action of a small perturbation. This is not
the case for a nonevolutionary discontinuity. At such a discontinuity, the
system of boundary conditions, which follow from the conservation laws, does
not have the unique solution for the amplitudes of outgoing waves generated
by given incident waves. From a mathematical view point this means that the
number of unknown parameters (the amplitudes of the outgoing waves and the
discontinuity displacement) is incompatible with the number of independent
equations. Since a physical problem must have the unique solution, the
assumption that the perturbation of a nonevolutionary discontinuity is
infinitesimal leads to a contradiction. In fact, the infinitesimal
perturbation results in disintegration, i.e., finite variation of the initial
flow, or transformation into some other unsteady configuration.
The evolutionarity requirement gives additional restrictions on the flow
parameters at a shock, compared to the condition of the entropy increase. The
restrictions appear because the direction of wave propagation (toward the
discontinuity surface or away from it), and thus the number of the outgoing
waves, depends on the flow velocity. If the velocity is large enough then the
given wave may be carried down by the flow. Therefore, at an evolutionary
discontinuity, the flow velocity must be such that it provides the
compatibility of the boundary equations. This form of evolutionarity
condition was applied to MHD shock waves in Refs. \cite{mhd,s58}. As a
result, the fast and slow shocks are evolutionary, while the
TASWs are nonevolutionary.
This classical picture was challenged when Wu \cite{wu87} obtained a
stationary TASW in a numerical simulation. The existence of a stationary
numerical solution does not mean of course that the shock is stable with
respect to disintegration or transition into another unsteady flow. Wu
\cite{wu88} demonstrated that a TASW, which is subfast upstream and subslow
downstream, disintegrates under the action of a small Alfv\'{e}n perturbation
with a large enough characteristic time. Nevertheless, this numerical
result was interpreted as being in a contradiction with the principle of
evolutionarity and stimulated the efforts to modify or even disprove this
principle.
It was suggested that the free parameters that describe a nonunique structure
of a TASW \cite{wu90} or the amplitudes of strongly damping dissipative waves
\cite{hada} should be included in the number of unknown parameters when
solving the problem of evolutionarity. This would make the TASW evolutionary.
In both cases, however, the perturbation is confined within the shock
transition layer. Consequently, it does not enter into the boundary
conditions, which relate the quantities far enough from the transition layer,
and thus it does not contribute to the evolutionarity \cite{evol}.
Wu \cite{wu90} also argued that the TASW whose nonevolutionarity is
based on separation of Alfv\'{e}n perturbations from the remaining
perturbations \cite{s58} becomes evolutionary in the case of a nonplanar shock
structure because in this case the separation formally does not take place.
However, as shown by Markovskii \cite{evol}, the coupling of the
small-amplitude Alfv\'{e}n modes with a low enough frequency to the remaining
modes is weak (unless the shock is of the type close to one of the degenerate
types, Alfv\'{e}n discontinuity or switch shocks). Therefore the coupling
becomes essential only when the small perturbation generates large variation
of the flow, which is the same result as predicted by the principle of
evolutionarity.
There is one more finding that favors the nonexistence of stationary
TASWs. As discussed by Kantrowitz and Petschek \cite{kp}, the TASWs are
isolated solutions of Rankine-Hugoniot problem, which do not have neighboring
solutions corresponding to small deviations of boundary conditions. Wu and
Kennel \cite{wk} introduced a new class of trans-Alfv\'{e}nic
shock-like structures with noncoplanar boundary states. The thickness of
such a structure increases in the course of time, and it eventually evolves
to a large-amplitude Alfv\'{e}n wave. It was thus shown that neighboring to a
TASW are time-dependent configurations, which are not solutions of the
Rankine-Hugoniot problem. In addition, Falle and Komissarov \cite{fk}
recently considered stationary TASWs of all possible types and showed that
the shocks disintegrate if the boundary values deviate from their initial
values.
Strictly speaking, a TASW, in contrast with fast and slow shocks, becomes a
time-dependent shock-like structure once it is perturbed by a small-amplitude
Alfv\'{e}n wave because the Alfv\'{e}n wave violates the coplanarity
condition. This fact, already on its own, means that the TASW becomes
unsteady under the action of the small perturbation. However, the scenario
for its evolution depends on the initial configuration and on the nature of
the perturbation. After the disintegration, the magnetic field reversal given
at the initial nonevolutionary shock may be taken either by a secondary TASW
or by an Alfv\'{e}n discontinuity. Both structures are nonevolutionary
\cite{m98,evol}. Therefore single disintegration does not lift the
contradiction inherent in a TASW. The main question that we solve in this
paper is what happens to the post-disintegration nonevolutionary
configuration. We show that the secondary TASW is again unstable with respect
to disintegration and that the evolution of a TASW may have the form of
oscillatory disintegration. In Sec. II, we describe the simulation method.
In Sec. III, we discuss the results of the calculations. Our conclusions are
presented in Sec. IV.
\section{Numerical method}
We take the MHD equations in the following form
\begin{mathletters}
\label{1}
\begin{equation}
{\partial \rho \over \partial t} + {\partial \rho v_{x} \over \partial x}
= 0,
\label{1a}
\end{equation}
\begin{equation}
{\partial \rho v_{x} \over \partial t} + {\partial \over \partial x}
\biggl( p + \rho v_{x}^{2} + {\textstyle {1\over 2}}{\bf B}_{\perp}^{2} -
{\textstyle { 4\over 3}} \eta {\partial v_{x} \over \partial x}\biggr)
= 0,
\label{1b}
\end{equation}
\begin{equation}
{\partial \rho {\bf v}_{\perp} \over \partial t} + {\partial \over
\partial x} \biggl( \rho v_{x} {\bf v}_{\perp} - B_{x} {\bf B}_{\perp}
- \eta {\partial {\bf v}_{\perp} \over \partial x} \biggr) = 0,
\label{1c}
\end{equation}
\begin{equation}
{\partial {\bf B}_{\perp} \over \partial t} + {\partial \over \partial
x} \biggl( v_{x} {\bf B}_{\perp} - B_{x} {\bf v}_{\perp}
- \nu_{m} {\partial {\bf B}_{\perp} \over \partial x} \biggr) = 0,
\label{1d}
\end{equation}
\begin{eqnarray}
&& {\partial \over \partial t} \biggl( {\textstyle {1\over 2}} \rho v^{2}
+ { p \over \gamma - 1 } + {\textstyle {1\over 2}} {\bf B}_{\perp}^{2}
\biggr) + {\partial \over \partial x} \biggl[ \rho v_{x}
\biggl({\textstyle {1\over 2}} v^{2}
\nonumber \\
&& \quad + { \gamma
\over \gamma - 1} { p\over \rho} \biggr)
+ \biggl( {\bf B}_{\perp} \cdot \biggl( v_{x} {\bf B}_{\perp} - B_{x} {\bf
v}_{\perp}
- \nu_{m} {\partial {\bf B}_{\perp} \over \partial x} \biggr)
\biggr)
\nonumber \\
&& \quad - \eta \biggl( {\textstyle{4\over3} }
v_{x} {\partial v_{x}\over \partial x}
+ \biggl( {\bf v}_{\perp} \cdot
{\partial {\bf v}_{\perp}\over \partial x}
\biggr)\biggr)\biggr] = 0.
\label{1e}
\end{eqnarray}
\end{mathletters}
Here the subscript "$\perp$" denotes the vector component perpendicular to
the $x$ axis, $B_{x}={\rm const},$ magnetic diffusivity $\nu_{m}$ and
viscosity $\eta$ are put constant and equal to 0.1 in all calculations, and
we use the units such that the factor $4\pi$ does not appear. The initial
distribution of the MHD quantities is given by the following formulas
\begin{mathletters}
\label{2}
\begin{equation}
\rho = { \textstyle {1\over 2}} (\rho_{\uparrow} + \rho_{\downarrow}) -
{ \textstyle {1\over 2}} (\rho_{\uparrow} - \rho_{\downarrow}) {\rm tanh}
( { x/ L}) ,
\label{2a}
\end{equation}
\begin{equation}
v_{x} = { \textstyle {1\over 2}} (v_{x\uparrow} + v_{x\downarrow}) -
{\textstyle {1\over 2}} (v_{x\uparrow} - v_{x\downarrow}) {\rm tanh} ( {
x/ L} ) ,
\label{2b}
\end{equation}
\begin{equation}
p = {\textstyle {1\over 2}} (p_{\uparrow} + p_{\downarrow}) -
{ \textstyle {1\over 2}} (p_{\uparrow} - p_{\downarrow}) {\rm tanh} ( {
x/ L}) ,
\label{2c}
\end{equation}
\begin{equation}
B_{y} = B_{\tau} {\rm cos} (\theta), \qquad
B_{z} = B_{\tau} {\rm sin} (\theta),
\label{2d}
\end{equation}
\begin{equation}
v_{y} = v_{\tau} {\rm cos} (\theta), \qquad
v_{z} = v_{\tau} {\rm sin} (\theta),
\label{2e}
\end{equation}
\begin{eqnarray}
B_{\tau} = && {\textstyle {1\over 2}} (\mid B_{\perp\uparrow}\mid + \mid
B_{\perp\downarrow}\mid )
\nonumber \\
&& - {\textstyle {1\over 2}}
(\mid B_{\perp\uparrow} \mid
- \mid B_{\perp\downarrow}\mid ) {\rm tanh} ( { x/ L} ),
\label{2f}
\end{eqnarray}
\begin{eqnarray}
v_{\tau} = && {\textstyle {1\over 2}} (\mid v_{\perp\uparrow}\mid + \mid
v_{\perp\downarrow}\mid )
\nonumber \\
&& - {\textstyle {1\over 2} }
(\mid v_{\perp\uparrow} \mid - \mid
v_{\perp\downarrow}\mid ) {\rm tanh} ( { x/ L} ),
\label{2g}
\end{eqnarray}
\begin{equation}
\theta = {\textstyle {\pi\over 2}} (1 + {\rm tanh} ( { x/ L})),
\label{2h}
\end{equation}
\end{mathletters}
where the subscripts "$\uparrow$" and "$\downarrow$" denote the quantities in
the asymptotic upstream and downstream regions, respectively.
After the
configuration relaxes to a steady state, it is perturbed by an Alfv\'{e}n
wave specified by the expression
\begin{mathletters}
\label{3}
\begin{equation}
B_{z} = {\textstyle {1 \over 2}} \delta B_{z} \biggl( 1 + {\rm tanh} \biggl(
{x - x_{0} \over l} \biggr) \biggr) ,
\label{3a}
\end{equation}
\begin{equation}
v_{z} = - B_{z} / \sqrt{\rho}.
\label{3b}
\end{equation}
\end{mathletters}
This wave moves downstream. The configuration is set by putting
$B_{x}=0.89,$ $L=1.45,$ and
\begin{mathletters}
\label{4}
\begin{equation}
B_{y\uparrow} = 0.93, \qquad B_{y\downarrow} = -0.8,
\label{4a}
\end{equation}
\begin{equation}
B_{z\uparrow} = v_{z\uparrow} = 0., \qquad
B_{z\downarrow} = v_{z\downarrow} = 0.,
\label{4b}
\end{equation}
\begin{equation}
v_{x\uparrow} = 1., \qquad v_{x\downarrow} = 0.55042,
\label{4c}
\end{equation}
\begin{equation}
v_{y\uparrow} = 1.04494, \qquad v_{y\downarrow} = -0.49476,
\label{4d}
\end{equation}
\begin{equation}
\rho_{\uparrow} = 1., \qquad \rho_{\downarrow} = 1.81681,
\label{4e}
\end{equation}
\begin{equation}
p_{\uparrow} = 0.00116, \qquad p_{\downarrow} = 0.56319.
\label{4f}
\end{equation}
\end{mathletters}
This corresponds to a ${\rm II \rightarrow III}$
shock, for which $V_{+\uparrow} >
v_{x\uparrow} > V_{Ax\uparrow}$ and $V_{Ax\downarrow} > v_{x\downarrow} >
V_{-\downarrow},$ where $V_{+}$ and $V_{-}$ are the fast
and slow magnetosonic velocities.
We solve Eq. (\ref{1}) using a uniform grid and an explicit conservative
Lax-Wendroff finite-difference scheme with physical dissipation \cite{pt}.
The time step is limited by the Courant-Friedrichs-Lewy (CFL) condition and
by the dissipation timescale. The boundary values are obtained by hyperbolic
interpolation. The numerical interval $-50 < x < +300$ is covered by 2600
grid points. The interval is chosen in such a way that no
large-amplitude wave reaches the boundaries during the computation time.
Small-amplitude waves pass through the boundaries without any detectable
reflection which could affect the flow inside the simulation region. We have
tested our code for a smaller mesh and a corresponding time step determined
by the CFL condition as well as for the same mesh and a time step smaller
than that determined by the CFL condition. The test showed that there is no
considerable dependence of our results on the mesh size and
time step.
\section{Results of simulations}
Equation (\ref{2}) does not exactly
describe the shock structure. Therefore the flow undergoes time variations
until it adjusts to a stationary shock transition. The resulting boundary
values differ slightly from those given by
Eq. (\ref{4}) but the difference is less than 1\%. The conservation laws for
these new values are fulfilled with the precision less than 0.1\%. The
stationary configuration is then perturbed by a small-amplitude Alfv\'{e}n
wave with $l=-L,$ $x_{0}= -40,$ and $\delta B_{z}= 0.025$ or $\delta B_{z}=
-0.025$ (Fig.~\ref{f1}). Note that $\delta B_{z}$ is about 50 times smaller
than $\mid {\bf B}_{\uparrow} \mid.$ Although in the case of an upstream
incident wave the perturbation of $B_{z}$ and $v_{z}$ (not shown) is carried
to the downstream region, the boundary conditions for the Alfv\'{e}n waves
are incompatible. Therefore the given Alfv\'{e}n perturbation pumps $B_{z}$
and $v_{z}$ into the shock or out of the
shock, depending on the sign of $B_{z}$ inside the transition layer. Since
$B_{z}$ inside the transition layer is nonzero, the shock behaves in
different ways under the action of the perturbations with positive and
negative $\delta B_{z}.$ If the shock and the perturbation carry $B_{z}$ of
the same sign, the shock disintegrates into a ${\rm II \rightarrow III}$
shock of a smaller amplitude, a large-amplitude slow shock,
and some other structures of a much smaller amplitude (Fig.~\ref{f2}a,b).
If the shock and the perturbation carry $B_{z}$ of opposite signs, the
situation is somewhat peculiar. The main secondary structures are a TASW and
a slow rarefaction (Fig.~\ref{f3}a,b). However, these structures do not
become separated. The reason is that the secondary TASW is of a so-called
${\rm II \rightarrow IV=III}$ type \cite{kbw}. This means that the downstream
velocity at the shock is exactly equal to the slow magnetosonic velocity.
Therefore there is no disintegration in the usual sense but the
configuration becomes unsteady because the right boundary of the slow
rarefaction moves away from the TASW, while the left
boundary remains attached to the shock. Note that the rarefaction wave is
attached to the TASW not at the density peak but somewhere to the right of
the peak. This is related to the fact that the density profile of a ${\rm II
\rightarrow IV}$ shock has a maximum (see, e.g., Ref. \cite{wu88}), in
contrast with the monotonic profile of a ${\rm II \rightarrow III}$ shock.
From the moment when the Alfv\'{e}n wave with $\delta B_{z}>0$
arrives to the shock, the disintegration starts almost immediately, in
contrast with the result of Wu \cite{wu88}. The reason is that the
disintegration time depends on the shock type and on its initial state. This
can be understood as follows. The important characteristic of a TASW,
introduced by Kennel {\it et al.} \cite{kbw}, is the integral of $B_{z}$ over
the
transition layer,
\begin{equation}
I_{z} = \int \limits_{x\downarrow}^{x\uparrow} B_{z} dx.
\label{5}
\end{equation}
This integral fixes the nonunique structure of a TASW. For a ${\rm II
\rightarrow III}$ shock, the quantity $I_{z}$ takes two distinct values,
$I_{z0}$ and $-I_{z0},$ and, for a ${\rm I \rightarrow III}$ or ${\rm II
\rightarrow IV}$ shock, it falls into the interval $-I_{z0} < I_{z} <
I_{z0}.$ The quantity $I_{z0}$ depends on the boundary values, and it tends
to infinity when the shock approaches an Alfv\'{e}n discontinuity or a switch
shock, which is intermediate between evolutionary and nonevolutionary shocks.
This result was obtained for almost parallel small-amplitude shocks, but one
may expect that it remains qualitatively valid in the general case.
When an Alfv\'{e}n wave is incident on a TASW, it changes $I_{z}.$ If we
start from a planar ${\rm I \rightarrow III}$ or ${\rm II \rightarrow IV}$
shock ($I_{z}=0$), as in the case studied by Wu \cite{wu88}, the quantity
$\mid I_{z} \mid $ first has to reach the value $I_{z0}.$ Only after that
it falls into the forbidden region, and the disintegration starts. In the
case of a ${\rm II \rightarrow III}$ shock, there is a different situation.
Since $I_{z}$ takes only distinct values $I_{z0}$ and $-I_{z0},$ the
disintegration starts immediately, and the disintegration time is close to
its minimum value $L/V,$ approximately equal to 30 in our case, where $V$ is
a relative velocity of the secondary discontinuities.
Let us now follow the further evolution of the post-disintegration
configuration under the action of a small perturbation. Our main conclusion
is that the secondary TASW is again unstable with respect to disintegration.
At the same time, the way of evolution depends on a form of the
perturbation. We first discuss the case where the perturbation of the
secondary TASW is such that $I_{z}$ continues to increase or
decrease, in particular where the perturbation is equal to its initial
positive (Fig.~\ref{f2}b,c) or negative (Fig.~\ref{f3}b,c) value. If the
perturbation of $B_{z}$ is positive then the shock spreads in space, with all
the jumps, except for $\Delta B_{y}$ and $\Delta v_{y},$ decreasing in time.
It thus approaches a large-amplitude Alfv\'{e}n wave. If the perturbation is
negative, the shock first passes through the state in which $I_{z}=0.$
This is not in a contradiction with the analytical theory, because a ${\rm II
\rightarrow IV}$ shock may have a planar structure, in contrast with a ${\rm
II \rightarrow III}$ shock \cite{kbw}. When $\mid I_{z} \mid$ reaches a
critical value, the shock disintegrates (Fig.~\ref{f3}b), and after that it
spreads in space approaching a large-amplitude Alfv\'{e}n wave
(Fig.~\ref{f3}c). The precursor of the disintegration is the peak in $B_{y}$
curve in Fig. \ref{3}b.
We now turn to a cyclic perturbation. We impose the perturbation described by
Eq. (\ref{3}) in such a way that $B_{z}$ changes sign at $x=-40$ and the
Alfv\'{e}n wave now carries the perturbation of the same amplitude but
opposite sign. After the first disintegration starts (at $t=20$ for
$\delta B_{z} >0$ and at $t=670$ for $\delta B_{z}<0$), the opposite sign
perturbation arrives to the TASW each 150 units of time. The resulting
configuration is such that the increase of $\mid I_{z} \mid$ is repeatedly
replaced by its decrease, and the shock undergoes oscillatory disintegration.
The disintegration configurations after several cycles are shown in
Figs.~\ref{f4} and \ref{f5}. As can be seen from the figures, the
configuration emits a sequence of contact discontinuities. The contact
discontinuities move with the flow velocity, which is approximately equal to
that given by Eq. (\ref{4c}). The corresponding time interval between the
discontinuities is equal to 150.
Downstream of the TASW, there is a wave train, which consists of slow shock
and rarefaction waves. These structures are not standing in the flow. They
consecutively emerge at the left edge of the train and merge at the right
edge. The merging is seen in Fig. \ref{4}b at $x\approx 150.$ We note that,
in the case of a negative initial perturbation, the transition through the
state with $I_{z}=0$ is not necessary for the oscillatory
disintegration to occur. If the perturbation changes sign for the first time
before $I_{z}$ becomes negative, the disintegration configuration is similar
to that shown in Fig. \ref{5}, except for the sign of $I_{z}$ inside the
shock.
Finally, the shock comes to a steady state only in a degenerate case where
the perturbation of the secondary TASW exactly compensates the nonzero value
of $B_{z}$ and $v_{z}$ outside of the transition layer. We emphasize that in
all but the degenerate cases the small Alfv\'{e}n perturbation makes the TASW
unsteady, in contrast with fast and slow shocks. However, there remains a
question. Formally, the initial TASW becomes a time-dependent structure, much
like the secondary TASW, since the Alfv\'{e}n perturbation arrives to the
initial shock. The question is why the initial TASW disintegrates when
$I_{z}$ increases monotonically, while the secondary TASW does not. To
answer this question, we first mention that the secondary TASW is more
close to a finite-amplitude Alfv\'{e}n wave than the initial shock.
Alfv\'{e}n waves, as well as switch shocks, are singular structures. As shown
by Kennel {\it et al.} \cite{kbw}, the quantity $dI_{z0}/dq$ tends to
infinity as the shock approaches these singular structures. Here
$q=B_{y\uparrow}/B_{y\downarrow}$ characterizes the jumps of the boundary
values at the shock with a given $I_{z},$ and $I_{z0}(q)$ is an allowed
curve in which a ${\rm II \rightarrow III}$ shock has a stationary structure.
Assume now that the initial shock is in the state $I_{z}=I_{z0}(q_{0}).$ A
small Alfv\'{e}n perturbation changes $I_{z}.$ For the shock to remain in the
curve $I_{z0}(q),$ a change of $q$ is required. In the general case, the
variation of $I_{z}$ is comparable with the variation of $q,$ and thus with
the jumps of the boundary values at the TASW. In this case, the evolution has
the form of disintegration. By contrast, if the shock is close to the
singular structure, the given variation of $I_{z}$ requires a small variation
of $q,$ and the jumps of the boundary values are adjusted to $I_{z0}(q)$ in a
diffusion-like manner. It should be mentioned that the curves $I_{z0}(q)$
were obtained by Kennel {\it et al.} \cite{kbw} for small-amplitude shocks
propagating almost parallel to the magnetic field. Nevertheless, we speculate
that, in our simulation, the initial TASW has a small enough
$dI_{z0}/dq$ to disintegrate, while for the secondary TASW the quantity
$dI_{z0}/dq$ is large enough to dim the disintegration. Such an explanation
does not imply that a TASW cannot disintegrate more than one time in
principle. Furthermore, in our simulation run with a positive constant
$\delta B_{z},$ there is an evidence for a possible second disintegration at
$t=380.$ However, the second disintegration is too faint to contend that it
indeed takes place.
\section{Conclusions}
We have performed a numerical simulation of a trans-Alfv\'{e}nic shock wave.
The shock that we have considered is of a ${\rm II \rightarrow III}$ type,
i.e., it is subfast upstream and superslow downstream. We have shown that the
shock disintegrates under the action of a small Alfv\'{e}n perturbation. The
resulting configuration includes a secondary TASW, a large-amplitude slow
shock or rarefaction wave, and other small-amplitude structures. We have also
demonstrated that the secondary TASW is again unstable with respect to
disintegration. When the perturbation has a cyclic nature, the shock
undergoes an oscillatory disintegration. This result is in a qualitative
agreement with our previous finding \cite{m98}. This process shows up
as a train of slow shock and rarefaction waves, which
consecutively emerge at one edge of the train and merge at the other edge. At
the same time, the disintegration configuration of a small-amplitude almost
parallel TASW discussed by Markovskii \cite{m98} includes alternating TASWs
and Alfv\'{e}n discontinuities rather than alternating TASWs. This
discrepancy is explained by the fact that, in the approximation used in Ref.
\cite{m98}, the difference between the secondary TASW and the Alfv\'{e}n
discontinuity manifests itself in higher orders.
In contrast with the results of Wu \cite{wu88}, the disintegration starts
almost immediately after the Alfv\'{e}n perturbation arrives to the
initial shock. The characteristic time of this process is equal to that
required for the secondary structures to become separated. The reason for
this can be seen as follows. TASWs have a nonunique structure. A ${\rm II
\rightarrow IV}$ shock transition studied by Wu \cite{wu88}, as well as a
${\rm I \rightarrow III}$ transition, allows a continuous family of integral
curves, while the ${\rm II \rightarrow III}$ shock has two distinct integral
curves. For given boundary values, each integral curve is fixed by the
definite parameter. The incident Alfv\'{e}n wave changes the parameter and
thus the shock structure. In the case of a ${\rm I \rightarrow III}$ or ${\rm
II \rightarrow IV}$ shock, some time passes until the parameter falls into a
forbidden region, and only after that the shock disintegrates. In the case of
a ${\rm II \rightarrow III}$ shock, its structure immediately becomes
inconsistent with the boundary values under the action of the Alfv\'{e}n
wave, which initiates the disintegration.
Thus, our simulations confirm that a TASW becomes unsteady when it is
perturbed by a small-amplitude incident wave. Furthermore, an almost
vanishing perturbation results in considerable dynamics at relatively small
timescales. The scenario for the shock evolution depends on its initial state
and on the nature of the perturbation. In particular, the evolution may have
the form of oscillatory disintegration in which the shock repeatedly
transforms into another TASW.
\acknowledgments{This work is supported in part by Russian Foundation for
Basic Research (grants 99-02-16344 and 98-01-00501).}
|
\section{INTRODUCTION}
Not long ago a variety of complex magnetic structures formed by many
strongly curved and entangled vortices was discovered in bulk
superconductors [1]. The origin of these structures can not be explained if
treat the motion of vortices like that of stick-like objects.
It is necessary to consider the evolution of three-dimensional magnetic
flux lines with potentially arbitrary shape.
For the first, it is useful to investigate the motion and shaping of a
single vortex but interacting with a surface supercurrent which represents
either transport current or Meissner current induced by external field. As
far as we know, even this simple task previously was not under careful
consideration.
Of course, a vortex never lives alone, without interactions with other
vortices, and no stable many-vortex structure could exist without mutual
repulsion of vortices. However, one can suppose that the scenario of
magnetic flux penetration into a current-carrying bulk superconductor should
be dominated only by vortex interaction with the surface current distribution,
i.e. eventually by geometry of superconducting sample, not by the
vortex-vortex interaction. The latter can not seriously affect this scenario,
merely because it itself is unable to ensure a deep penetration at all.
To prove this statement, let us imagine the steady flow of vortices which
arise at the flat current-carrying boundary of half-infinite superconductor
and then move deep into the sample due to their repulsion. Clearly, because
of viscous character of vortex motion, such the flow needs in nonzero
gradient of concentration of vortices. As a consequence, both the
concentration and the local drift velocity of vortices must be decreasing
functions of the depth. Hence, their product is not constant, that is the
magnetic flux conservation can not be satisfied. This discrepancy means that
no steady flow could be supported by the inter-vortex forces only. In
particular, it is impossible to realize the stationary lasting penetration
of vortices from infinite flat boundary parallel to external magnetic field.
Therefore, the only force what can push a vortex through the sample interior
is nothing but self-action of vortex caused by its distorsion. But in order
to involve this force into the evolution, the vortex must feel the shape of
the sample boundary.
Hence, the true picture looks as follows.
After nucleation in a surface layer with thickness of order of
London penetration depth $\lambda $ , a vortex firstly expands over the
sample boundary remaining in this layer. At this stage only the end
fragments of vortex are factually moving. The ends slide along the boundary,
and the resulting shape of vortex core reflects that of boundary. This process
lasts until the curvature of main middle part of the core becomes
sufficiently strong in order to cause the deepening of vortex as a whole.
In view of these reasonings, the geometry of steady transport of magnetic
flux into a bulk supeconductor looks rather insensible to inter-vortex
interactions and thus can be testified in terms of a single vortex, at
least if not consider details of vortex nucleation and processes like
annihilation and reconnections of vortices [1] which take place deep inside
the sample.
For example, many aspects of resistivity in supercurrent-carrying wires can
be described as evolution of ring-like vortices as if thats instantly arise
near the boundary, then contract independently one on another and finally
self-annihilate [2]. However, more correct consideration should include the
first stage when vortex transforms from small nucleus into a ring. We shall
see that in fact this stage may result also in a non-ring penetration geometry,
and more detailed theory can predict what the scenario realizes
under given transport current value and sample dimensions.
Though a lot of works were published previously touching upon a role of
vortex distorsions, for instance, under a pinning by randomly distributed
centers [3], always some preliminary restrictions of the vortex
geometry were attracted. In the present work the general equations of
evolution of arbitrarily curved vortex lines in isotropic superconductors
are formulated and analysed. We shall especially discuss the true
formulation of boundary conditions for these equations.
It will be shown that in a sample whose dimensions noticably exceed
$\lambda $ the vortex can penetrate either as flexible stick or as elastic
(similarly to a rubber thread attracted by its ends through water).
of surface supercurrent. The latter case occurs only if surface
supercurrent exceeds $H_{c1}c/4\pi $ (in CGS units) and
is characterized by giant stretching of the vortex core along the
sample boundary in the direction parallel to drift of the ends.
The stretching is accompanied by decrease of both the vortex
energy and viscous dissipation per unit drift velocity, and
results in strong increase of the vortex drift velocity under given
transport current. In the framework of this scenario, the vortex core
firstly tranforms into a ring-like curve winding round the wire
cross-section (or into a spiral, if there is an external magnetic
field parallel to current), and only later the vortex begins to cut
the wire and enter deep into its interior.
This general picture is in agreement with the known simplific model
of magnetic flux penetration into round wires. Additionally,
our approach allows to scope very different stages of vortex evolution in
unified manner and obtain quantitative estimates for each stage.
\section{LONDON APPROXIMATION}
We shall confine ourselves by the London approximation. Of course, it would
have no sense if one could not apply it to actually moving vortices. But in
any case the requirement must be satisfied that characteristic velocity $u_0$
of viscous vortex motion influenced by magnetic fields comparable with
low critical magnetic field $H_{c1}$ , must be significantly smaller than
the speed of electromagnetic waves.
The velocity scale $u_0$ can be naturally estimated as
\[
u_0\equiv \mu \varepsilon /\lambda
\]
where
\[
\varepsilon =\Phi _0H_{c1}/4\pi
\]
is the self-energy per unit length of long straight-line vortex,
$\varepsilon /\lambda $ is the characteristic scale of Lorentz force also
related to unit length, and $\mu $ is mobility of the vortex core.
Below it will be seen that so defined $u_0$ really serves as the velocity
unit. If combine this definition and the known relations [4]
\[
\frac{c^2}{\Phi _0\mu }\sim \sigma _nH_{c2}\,\,\,,\,\,\,\sigma _n\sim
\frac \hbar \Delta (\frac c\lambda )^2
\]
with standard notations, $\sigma _n$ being the normal conductivity and
$\Delta \sim 2k_BT_c$ being the order parameter, one obtains
\[
u_0\sim \frac{\lambda k_BT_c}{2\pi \hbar }\frac{H_{c1}}{H_{c2}}
\]
As a typical example, at
$T_c\sim 100\,K$ , $\lambda \sim 3\cdot 10^{-5}\,cm$ and
$H_{{c2}}/H_{c1}\sim 500$ , one gets the estimate $u_0\sim 10^5\,cm/s$ .
This value looks small enough to allow for applicability
of quazi-static London approximation.
In fact, such an approach was used in large number of works on motion of
separate vortices as well as vortex lattices. The obvious exception is very
dense lattice, with small inter-vortex distancies of order of coherence
length. But our present subjects of interest are far from such complications.
\section{EVOLUTION EQUATIONS}
In the framework of London approximation, the free energy $E$ of vortex,
placed into a given surroundings, is completely determined by the shape of
its core, $R(p)=\{X(p),Y(p),Z(p)\}$ , with $X,Y,Z$ being coordinates of
the core points and $p$ being a scalar parameter. In accordance with the
principles of mechanics and nonequilibrium thermodynamics, the simplest
equation of a massless viscous evolution of the core line looks as
\begin{equation}
\mu^{-1} \partial R/\partial t= f(R)
\end{equation}
with Lorentz force on the right-hand side and friction force on the
left, both being related to unit core length.
By its sense, the parameter $\mu ^{-1}$ is the effective drag coefficient
which is determined by all the dissipative energy losses conjugated with the
core motion. Generally, there are at least two sorts of dissipative
processes accompanying the motion (see, for example, the review [4]),
namely, relaxation of the order parameter and normal currents
induced by time-dependent own magnetic field of the vortex. A concrete
expression for $\mu $ can be derived from more detailed theory, for
instance, from the Ginzburg-Landau functional approach, under its reduction
to London approximation [4]. The reduction is possible because normal
self-current of moving vortex and corresponding dissipation are located
mainly in a close vicinity of the core line, at distance comparable with
coherence length. After the transition to London's description, the effect
of normal currents becomes hidden in $\mu $ , but these currents give no
contribution to the Lorentz force [4]. Therefore, the reduction results in
the identity whose meaning is balance of friction force and Lorentz
force, as it is stated by the Eq.1, with $f(R)$ being determined only
by supercurrents.
To write $f(R)$ , one has not to evaluate the supercurrent
distribution. Instead, as in general in mechanics and statistical
thermodynamics, $f(R)$ can be expressed by means of $E$'s variation under a
small displacement of a local core fragment, that is as the functional
derivative $\delta E/\delta R(p)$ . However, the latter itself is not
invariant with respect to arbitrary (non-degenerated) transformations of the
parametrization $R(p)$ and to physical dimensionality of $p$ . In case of
isotropic media, the only true invariant expression for the Lorentz force is
\begin{equation}
f=-\frac{dp}{dL}\frac{\delta E}{\delta R(p)}=-\left|
\frac{\partial R}{\partial p}\right| ^{-1}
\frac{\delta E}{\delta R(p)}=\frac{\Phi _0}c\left[ J\times N\right]
\end{equation}
Here the vector $\partial R/\partial p\equiv R^{\prime }$ is locally
parallel to the core, $dL=|R^{\prime }|\,dp$ is the differential of the core
length,
$N\equiv R^{\prime }\left| R^{\prime }\right| ^{-1}=\partial R/\partial L$ ,
and $J$ is the density of full effective supercurrent which streams around
core and pushes a given core fragment.
The energy $E=E\{R(p)\}$ includes self-interaction of vortex and
its interaction with surroundings, in particular, with other vortices.
Correspondingly, in general $J$ consists of external currents and
self-current of vortex determined by its distorsion.
The Eqs.1 and 2 could be directly extended to a number of interacting
vortices. Besides, in principle, one may add
into $E$ also interactions with pinning potentials. However, below we
are interested only in motion of separate vortex in absence of pinning.
The parameter $p$ enumerates strictly the core points. But in practice it is
preferable to use another kind of parametrization, concretely, to introduce
the parameter $q$ which enumerates some suitable continuum of surfaces
$Q(r)=q$ , $r=\{x,y,z\}$ , each possessing only one intersection with
core line. The connection between $p$ and new parameter $q$ is
implied by the obvious relation $Q(R(p(q,t),t))=q$ , and simple algebraic
manipulations lead to the modified form of the evolution equations,
\begin{equation}
\frac{\partial R}{\partial t}=\mu
\left[ 1-\frac{\partial R}{\partial q} \otimes
\frac{\partial Q(R)}{\partial R}\right]f(R)\,\,,\,\,\,
\,f(R)=-\left| \frac{\partial R}{\partial q}
\right| ^{-1}\frac{\delta E}{\delta R(q)}
\end{equation}
where the symbol $\otimes $ denotes the tensor product of two vectors.
These equations describe how the intersection points marked by $q$ move
along the corresponding surfaces $Q(r)=q$ . Clearly, this is factually
two-dimensional motion. This feature becomes quite obvious
if it is possible
to identify $q$ as one of cartezian coordinates, that is to use parallel
planes as the marking surfaces. For instance, if thats are XY-planes, $q=Z$
and $Q(r)=z$ , then the Eqs.3 reduces to the equation
\begin{equation}
\frac \partial {dt}\left(
\begin{array}{c}
X \\
Y
\end{array}
\right) =-\frac \mu {\sqrt{1+X^{\prime 2}+Y^{\prime 2}}}\left(
\begin{array}{cc}
1+X^{\prime 2} & X^{\prime }Y^{\prime } \\
Y^{\prime }X^{\prime } & 1+Y^{\prime 2}
\end{array}
\right) \left(
\begin{array}{c}
\delta E/\delta X(Z) \\
\delta E/\delta Y(Z)
\end{array}
\right)
\end{equation}
with shortened notations $X^{\prime }\equiv \partial X/\partial Z$ ,
$Y^{\prime }\equiv \partial Y/\partial Z$ . The Eq.4 describes the time
evolution of $X$ and $Y$ coordinates of the core points marked with
their $Z$-coordinate.
\section{BOUNDARY CONDITIONS AND LOCAL APPROXIMATION}
In absence of pinning and more vortices, the vortex energy $E=E_s+E_i$
consists of two parts: the energy $E_i$ of the vortex interaction with
transport or Meissner supercurent and the self-energy $E_s$. Therefore, the
current in the Eq.2 also can be devided into two parts, $J=J_s+J_i$ .
Formally, $E_s$ is a complicated spatially non-local functional [5]
depending on both the core configuration $R(p)$ and the shape of sample.
Among other factors, $E_s$ includes the vortex interaction with the sample
boundary what can be interpreted as attraction of the end fragments of the
core to their mirror images placed outside superconductor.
But, if the curvature radius of the core everywhere is not too small as
compared with $\lambda $ , and besides, if the core nowhere is too close to
itself, then the so-called local approximation is possible,
\begin{equation}
E_s\approx \varepsilon L=\varepsilon \int \left| dR(p)\right|
\end{equation}
where $L$ is the core length. This well known approximation was argued and
used as long ago as in 1968 by Galaiko [6], and later by many other authors
(in particular, in [2-5]). Our own computer simulations showed that the
relative error of evaluation of self-action force by means of local
approximation does not exceed a few percents even if the curvature radius is
as small as $0.1\lambda $ .
In the local approximation the Eqs.1 and 2 take the form
\begin{equation}
\frac{\partial R}{\partial t}=\mu \frac{\Phi _0}c\left[ (J_s+J_i)\times
N\right] \,\,\,\,\,,\,\,\,\,\,J_s=\frac{cH_{c1}}{4\pi }\left[
N\times \frac{\partial ^2R}{\partial p^2}\right] \left|
\frac{\partial R}{ \partial p}\right| ^{-2}
\end{equation}
Here $J_s$ is the self-current what flows through the very core. Its
absolute value is inversely proportional to the local curvature
radius of core.
However, the local approximation needs to be accompanied by correct boundary
conditions. The true conditions should take into account the vortex
interaction with superconductor boundary. There are two ways to show that
this interaction results in the orthogonality of the end fragments of core
to the boundary. Thought these conditions are known at least since [6],
sometimes thats are neglected, so it is desirable to present more
argumentation.
First, let us note that the force vector $f(R)$ is always perpendicular to
the local core direction. Indeed, any variarion $\delta R$ parallel to this
core direction, $\delta R\parallel \partial R/\partial p$ ,
merely is identical to a change of parametrization,
without factual change of the shape, so
it has no physical meaning and should result
in $\delta E=0$ (therefore the
last expression in (2) always is consistent with previous ones). The same is
seen from (6). As a consequense, any core point displaces perpendicularly to
the core, in particular, the end points do which are placed
just on the boundary.
Hence, we must conclude that the end fragments
always are oriented to be orthogonal with respect to the boundary.
Secondly, the non-orthogonality would mean that the contour formed by
core and its mirror image is broken at the end point, i.e. has infinitely
small curvature radius here. From the point of view of exact $E_s$ [5],
if such a sharp "knee" occured it
would cause infinitely strong Lorentz self-action force and
consequently would be immediately straightened thus restoring the
orthogonality.
But, we must to underline that the principal conclusions to be deduced do
not refer to the local approximation and can be derived from general
non-local Eqs.1 and 2 only.
\section{STICK-ELASTIC VORTEX TRANSFORMATION\\ IN CURRENT-CARRYING PLATE}
To avoid a complicated mathematics, we confine ourselves by
simplific superconductor geometry. Consider the vortex evolution in an
infinitely wide plate, $-D<Z<D$ , without pinning but in presence of
transport surface supercurrent uniformly distributed over the boundary
planes and obeying the London equation. If this current flows along Y-axis
then
\[
J_x=J_z=0\,\,\,\,\,,\,\,\,\,J_y=\frac c{4\pi \lambda
}H_{c1}j(Z)\,\,\,\,,\,\,\,j(Z)\equiv h\frac{\cosh (Z/\lambda )}{\cosh
(D/\lambda )}
\]
with $h$ being the dimensionless measure of current density.
Let initially the vortex pierces the plate in Z-direction being described
with $R=\{X(Z,0)=0,0,Z\}$ . It has similar orientation soon after nucleation
near the edge of a real finite plate. Then, due to obvious spatial symmetry,
the vortex will remain inside the XZ-plane $Y=0$ and keep only one
intersection with any of XY-planes. In this situation the Eq.4 can be
applied and, besides, reduced to only equation for
X-coordinate, $X=X(Z,t)$ , as a function of time and Z -coordinate:
\begin{equation}
\frac{\partial X}{\partial t}=-\mu
\sqrt{1+X^{\prime 2}}\frac{\delta E}{\delta X(Z)}
\end{equation}
The energy can be expressed as
\begin{equation}
E=E_s+E_i=E_s-\frac \varepsilon \lambda \int X(Z,t)j(Z)dZ
\end{equation}
where the integral represents
the energy $E_i$ of vortex interaction with transport
current (this expression differs only by some constant from the general $E_i$
representation [5]). In the local approximation (5), the Eq.7 looks as
strongly nonlinear diffusion-type equation
\begin{equation}
\frac{\partial X}{\partial t}=u_0\left[ \lambda
\frac{X^{\prime \prime }}{1+X^{\prime 2}}+\sqrt{1+X^{\prime 2}}j(Z)\right]
\end{equation}
with notation $X^{\prime \prime }\equiv \partial ^2X/\partial Z^2$ and
characteristic velocity $u_0$ introduced in Sec.2.
Here the left side is responsible for the friction, and two terms on the
right-hand side represent the self-action force and transport
current-induced Lorenz force, respectively. Clearly, because of the latter
force both the vortex ends will forwardly move in one and the same direction
(to opposite edge of the plate), while the middle of vortex will be more or
less backward, and the larger is transport current the longer should be the
distance $\Delta X = X(\pm D,t)-X(0,t) $ (below termed vortex stretching).
Some predictions of further vortex behaviour can be deduced merely from the
energy expression (8). Just after start the middle is still in rest. As the
Eq.8 shows, in thick plate ($D>>\lambda $) the unit displacement of every
end leads to the $E_i$ 's decrease by $h\epsilon $ . At the same time, the
corresponding lengthening of each of two symmetrical core branches results
in the $E_s$ 's increase by $\epsilon $ per unit length. Consequently, if
$h>1$ then the total energy decreases,
and the vortex stretchening along the
drift direction becomes profitable. The lengthening process should last
until the curvature of the most backward central part of the core becomes so
large that the self-action force makes this part moving as quickly as the
ends do.
Thus, at $h>1$ the vortex gets over the friction like elastic in water.
Oppositely, at $h<1$ the stretchening is energetically unprofitable, and the
vortex should move as deformed flexible stick. The transition from this
stick-like behaviour to elastic-like one, when
transport current increases from $h<1$ to $h>1$ ,
is the example of so-called "nonequilibrium phase transitions".
Let us consider the steady drift of vortex as a whole, without change of
shape. The corresponding solution on the Eqs. 7 or 9 can be written as
$X(Z,t)=ut+X(Z)$ . The stationary shape $X(Z)$ and the drift velocity
$u=u(h,D)$ should be obtained from (7) or (9) with the help of above
discussed orthogonality boundary conditions
\[
\frac{dX}{dZ}(\pm D)=0
\]
Besides, due to the mirror symmetry,
the condition $X^{\prime }(\pm 0)=0$ should be satisfied.
In this steady nonequilibrium state the self-energy $E_s$ is constant,
therefore, the work $M_j$ produced by transport current per unit time,
\[
M_j=-\frac{dE_i}{dt}=\frac{u\Phi _0}c\int_{-D}^DJ_y(Z)dZ=2uh\varepsilon
\tanh (D/\lambda )
\]
coinsides with the energy dissipation per unit time $M_d$ . In accordance
with (1) and (2),
\[
W_d=\frac 1\mu \int \left| \frac{\partial R}{\partial t}\right| ^2dL=\frac
1\mu \int \left( \frac uQ\right) ^2QdZ=2\delta u^2/\mu
\]
where the notations
\[
Q\equiv \sqrt{1+X^{\prime 2}}=\frac{dL}{dZ}\,\,\,,\,\,\delta \equiv
\int\limits_0^D\frac{dZ}Q\,
\]
are introduced. We took into account that actual displacement of the core
always is locally perpendicular to its orientation. Only such displacements
are physically meaningful and really cause the friction. Therefore, the
drift velocity and the local core velocity are connected by the relation
\[
\left| \frac{\partial R}{\partial t}\right| =\frac uQ
\]
Evidently, the factor $Q$ determines at one and the same time local
orientation of the core and degree of its stretching.
Hence, the equality $M_d=M_j$ yields
\begin{equation}
U\equiv \frac u{u_0}=\frac{h\lambda }\delta
\tanh (D/\lambda )\approx \frac{ h\lambda }\delta
\end{equation}
In view of above reasonings, at $h<1$ the vortex stretchening is weak,
therefore, $X^{\prime 2}$ is comparable with unit, $Q\sim 1$,
$\Delta X\sim D$ and $\delta \sim D$.
Then the Eq.10 shows that in this stick-like regime
$U\sim h\lambda /\delta \approx h\lambda /D<<h$ ,
i.e. the drift velocity is
inversely proportional to the plate thickness. This is quite natural,
because the surface current-induced Lorentz force acts only on the ends,
while the friction almost equally acts on any core fragment.
In general, the parameter $\delta $ serves as the effective plate
half-thickness. Obviously, always $\delta <D$ . In the stretched
elastic-like regime in thick plate anywhere at $D-|Z|>> \lambda $ the
inequalities $\left| X^{\prime }\right| >>1$ and $Q>>1$ take place.
Hence, $\delta <<D$ and what is essential it becomes almost
insensitive to thickness. As a consequence, both the drift velocity and
mobility $u/h$ strongly increase as compared with stick-like regime and both
become independent on thickness (below we shall see that $\delta \sim
\lambda $ and $U\sim h$ , i.e. $U$ becomes approximately $D/\lambda $ times
larger). According to the $M_d$'s expression, the matter is that thought
the energy dissipation $Q$ times increases due to the core lengthening this
effect is overpowered by its $Q^2$ decrease because of $Q$ times decrease of
the factual core velocity $\left| \frac{\partial R}{\partial t}\right| $ .
As the result, the vortex stretching leads to smaller friction and smaller
entropy production, under fixed vortex velocity, and to larger velocity
ander fixed transport current. The picture looks as if most part of core
slides along itself, but this process does not mean a real motion of core
and so does not cause a friction and dissipation.
\section{DRIFT OF THE VORTEX ENDS}
To be convinced in what was said, let us consider vortex shape in the
stretched elastic-like regime. Because at $\Delta X>>D$ most part of the
core inevitably has a small curvature, it can be considered with neglecting
the self-action. Then any of the Eqs.7 and 9 reduces to
\begin{equation}
U\approx \sqrt{1+X^{\prime 2}}j(Z)
\end{equation}
Here from the characteristic exponential asymptotics does follow,
\begin{equation}
X(Z)\approx -\frac{\lambda U}h\left[ \exp \left( \frac{D-Z}\lambda \right)
-1\right]
\end{equation}
Here $Z>0$ , $X(-Z)=X(Z)$ , and for definitness the position $X=0$ is
prescribed to the end fragments. It is easy to verify that corresponding
self-action contribution in the Eq.9 indeed is negligibly small as compared
with the current-induced force.
We can get a rough estimate of the stretching if put on $Z=0$ in (12) and
take into account that $U>h\lambda /D$. Then the Eq.12 yields
\[
\Delta X/\lambda \approx \frac Uh\exp \left( \frac D\lambda \right) >\frac
\lambda D\exp \left( \frac D\lambda \right)
\]
Hence, $\Delta X/\lambda $ possesses exponentially strong dependence on
$D/\lambda $ , and it can be giantly large if $D$ exceeds $\lambda $ by an
order of value or more.
In view of this circumstance, the ratio $\Delta X /W $ with $W$ being the
width of a real finite plate, becomes of principal importance. Clearly, if
$\Delta X >> W $ then the steady drift of the vortex as a whole is
impossible: the ends of vortex will achieve the opposite edge before the
displacement of its backward central part will be comparable with $W$ (all
the more, before the velocity of this part becomes equal to that of the
ends).
Consider the drift of the ends in such a non-stationary situation. Because
the vortex lengthening is profitable, this drift can do independently on the
motion of deepened backward part, as if thickness was infinitely large
($D/\lambda \rightarrow \infty $) .
To estimate the drift velocity, let us
multiply the Eq.7 or 9 (with $\frac{\partial X}{\partial t}\Rightarrow u$ )
by $Q^{-1}$ and integrate over variable $z=D-Z$ from zero to infinity,
with the condition $X^{\prime }(z\rightarrow \infty )=\infty $ which
evidently corresponds to infinitely far backward center. Then both the Eqs.7
and 9 result in
\[
U=\frac \lambda \delta (h-1)\,\,\,,\,\,\delta =\int_0^\infty \frac{dz}Q
\]
To evaluate this integral, note that in accordance with the orthogonality
boundary conditions the shape of the end fragments of the core is parabolic,
for instance, at upper end $X(Z)=X(D)-(Z-D)^2/2\rho $ , with $\rho $ being
the curvature radius at the end point. It follows from the Eq.9 that
$\lambda /\rho =h-U$ .
In this parabolic region the integration divergers
but becomes cut after transition to exponential asymptotics (12). The
estimate of the integral leads to approximate equation
\begin{equation}
U\approx h(h-1)/\{h-1+\ln [2z_0(h-U)/\lambda ]\}
\end{equation}
where $z_0$ is the depth of the crossover point, $z_0\sim 4\lambda $ . The
Eq.12 helps to estimate the end drift velocity in thick plate. Obviously, it
turns into zero at $h\rightarrow 0$, in agreement with $D\rightarrow \infty $
limit of the estimate for stick-like regime. It can be shown that velocity
of the steady drift of the vortex as a whole is only slightly smaller
differing by a multiplier of order of unit.
\section{GIANT VORTEX STRETCHING}
The exponentially large vortex stretching is the most significant
possibility of vortex evolution in thick plate, as well as in bulk samples
in general. Therefore it would be useful to more correctly justify the above
simplific estimate of $\Delta X $ .
Note that $\Delta X>L/2-D$ . Divide both sides of (9) by $j(Z)$ and
integrate from zero to $D$ . This results in
\[
L/2=B-A\,\,,\,\,B\equiv U\int \frac{dZ}{j(Z)}\,\,,\,\,\,\,A\equiv \int
\arctan (X^{\prime })\left| \frac d{dZ}\frac \lambda {j(Z)}\right| dZ
\]
It is easy to notice that
\[
A<A_0\equiv \frac{\pi \lambda }2\left[ \frac 1{j(0)}-\frac 1{j(D)}\right]
\]
so $\Delta X>B-A_0-D$. The calculation of integral $B$ gives
\[
B=\frac{2U\lambda }h\cosh (D/\lambda )\{\arctan [\exp (D/\lambda )]-\frac
\pi 4\}
\]
Then, after simplifications possible due to $D>>\lambda $, one finally
obtains
\begin{equation}
\Delta X/\lambda >\frac{\pi (U-k)}{4h}\exp (D/\lambda )\,
\end{equation}
with $k<1$ . Because $u$ is monotonously growing function of $h$ , the
coefficient in front of exponent is positive if $h$ exceeds some level
larger than unit, for example, if $h>2$ . Hence, ate least at $h>2$ the
vortex stretching is exponentially strong.
This estimate is obtained in the framework of local approximation. More
correct estimate should give a lesser value, because of self-attraction of
the core in middle part of the plate where two symmetrical exponential tails
described by (12) meet one another and form an arc. Such a non-local effect
is most essential just under the specific plate geometry. However, the
non-local correction can not change the shape of the front vortex part where
the non-local interaction is weak as compared with other forces. It is
not hard to show that the latter requirement is
satisfied if $|Z|>Z_0$, where $Z_0$ is the solution on
equation
\[
h\exp [-(D-\left| Z\right| )/\lambda ]\approx \sqrt{\lambda /2\pi \left|
Z\right| }\exp (-2\left| Z\right| /\lambda )
\]
If take into account that the more is $h$ the less is $Z_0$, then this
equation yields $Z_0<D/3$ . Hence, at $|Z|>D/3 $
the mutual attraction of two core branches can be neglected, and the
asymptotics (12) remains valid. This means that the maximally possible
effect of non-locality is the replacing $D$
in the exponent by $\alpha D$ with $\alpha >2/3$ . Consequently, the lower
bound for the stretching with confidence can be estimated as
\[
\Delta X/\lambda >\frac \lambda D\exp (2D/3\lambda )
\]
Thus, even in the worst case the non-local effects
do not abolish the exponential character of stretching.
For example, if $\lambda \approx 3\cdot 10^{-5}\,cm $
and $h$ equals to a few units, then
even at $D\sim 20\lambda <10^{-3}\,cm$ one gets
$\Delta X >m\lambda \exp(2D/3\lambda ) $ , with $m \sim 1 $ ,
i.e. $\Delta X > 1\,cm $ what exceeds a width of any realistic sample.
Thus at first the vortex should form a ring whose shape approximately copies
that of the sample cross-section. During this process the velocity of
backward deepened core part is primarity determined by its distorsion
which is created in the beginning of stretchening and thus has
curvature radius of order of $D$ . Hence, this velocity is of
order of $u_0\lambda /D$ , and at the moment when the ends will meet
one another the displacement of most backward point will be
yet as small as $\sim \lambda W/Dh<<W$ .
\section{DISCUSSION AND RESUME}
It seems clear that both the above conclusions can be extended to bulk
current-carrying superconductors with another geometry, for instance, to
round wires, if treat $2D$ and $W$ as minimal and
maximal diameters of cross-section of
the wire, respectively.
Due to possibility of giant deformation and stretching of vortices, the
thermodynamically nonequilibrium process of vortex penetration
can promote formation of complicated many-vortex dynamical configurations
which seem rather strange and unprofitable from the point of view of
equilibrium thermodynamics. The presence of an external magnetic field
parallel to transport current should lead to formation of
spiral-like configuration instead of ring-like one and thus especially
ensure the entangling of vortices. The Eq.4 enables us to describe
this scenario in details, if choose Z-axis to be directed along the wire.
Besides, the presence of weak pinning should amplifier the stretching
of vortex and additionly complicate its shaping,
because the motion of deepened part of
vortex is characterized by relatively small forces of order of
$\epsilon \lambda /D $ (much smaller than forces $\sim \epsilon $
what act on the end fragments) and so may be easily held
back by pinning centers.
We would like to underline the role of orthogonality boundary conditions.
In the work [3] the equation was under use similar to
our Eq.9, but boundary conditions was formulated in terms of the tension
of core line. One can see from [3] that such conditions make it
impossible to
consider the case of high surface transport current $>H_{c1} $
corresponding to the elastic-like regime.
To resume, we formulated the
invariant equations of viscous motion of arbitrarily shaped 3D vortex
lines, and applied them to careful analysis of the scenario of
vortex penetration into a thick superconducting sample. As it was
argued, the vortex-vortex interaction does not significantly affect
the penetration process. But, of course, a full description of
resistive state leads to more complicated tasks about vortex-vortex
interactions deep inside the sample.
\,\,\,\,
ACKNOWLEDGEMENTS
I would like to thank Dr. M.Indenbom, Dr. Yu.Genenko and Dr. A.Radievskiy
for usefull discussions.
\,\,\,\,
REFERENCES
1. M.V.Indenbom, C.J.van der Beek, V.Berseth, W.Benoit, G.D'Anna,
A.Erb, E.Walker and R.Flukiger, Nature, 1997, Feb.20 .
2. Yu.A.Genenko, Phys.Rev., B 49, 1994, 6950.
3. Chao Tang, Shechao Feng and L.Golubovich, Phys.Rev.Lett.,
472, 1994, 1264.
4. L.P.Gorkov and N.B.Kopnin, Uspekhi fizicheskikh nauk, 116, 1975, 411
(transl. in English in Sov.Phys.-Usp., 1975).
5. Yu.E.Kuzovlev, Physica, C 292, 1997, 117.
6. V.P.Galaiko, Zh.Teor.Eksp.Fiz., 50 , 1966, 1322.
\end{document}
|
\section{Introduction}
Hybrid baryons are bound states of three quarks with an explicit
excitation in the gluon field of QCD.
The construction of (hybrid) baryons in a
model motivated from the strong coupling expansion of the hamiltonian
formulation of lattice QCD, the non--relativistic
flux--tube model of Isgur and Paton \cite{paton85}, was detailed in ref. \cite{hadron}.
This model predicts the adiabatic
potentials of (hybrid) mesons at large interquark separations,
as well as the mass of the $J^{PC}=1^{-+}$ hybrid meson,
consistent with recent estimates from
lattice QCD \cite{paton85,morning}.
In ref. \cite{hadron} we studied the detailed flux dynamics and built
the flux hamiltonian. We restrict our discussion to cases where
the flux settles down in a Mercedez Benz configuration (as motivated by
lattice QCD \cite{bali}).
A minimal amount of
quark motion is allowed in response to flux motion,
in order to work in the centre of mass frame.
Otherwise, we make the so--called ``adiabatic'' approximation, where the
flux motion adjusts itself instantaneously to the motion of the quarks.
The main result is that the lowest
flux excitation can to a high degree of accuracy (about 5\%) be simulated
by neglecting all flux--tube motions except the vibration of a
junction. This result was obtained within the small oscillation approximation.
The junction acquires an effective mass $M_{\mbox{\small eff}}$ from the motion of the
remainder of the flux--tube and the quarks.
The model is then simple: a junction is
connected via a linear potential to the three quarks.
The ground state of the junction motion corresponds to a conventional
baryon and the various excited states to hybrid baryons.
The junction can move in three directions, and correspondingly be excited
in three ways, giving the hybrid baryons $H_1, H_2$ and $ H_3$.
The junction motion is depicted in Fig. \ref{rough}.
\begin{figure}[t]
\vspace{0cm}
\begin{centering}
\epsfig{file=captalkfig1.ps,width=15cm,angle=0}
\vspace{-0.2cm}
\caption[x]{The junction connects strings coming from the three quarks. The vectors
$\br$ and $\bl_i$ respectively point from the equilibrium position
of the junction to its current position and the quark positions.}
\plabel{rough}
\end{centering}
\end{figure}
\vspace{0.3cm}
The hamiltonian for the junction motion in the Mercedez Benz configuration
is simply the kinetic energy of the junction added to the sum of the lengths
from the junction to the quarks multiplied by the string tension $b$,
\begin{eqnarray} \plabel{ham3}
{H}_{\mbox{\small flux}}=\frac{1}{2}M_{\mbox{\small eff}}\; {\bf \dot{r}^2}+
b\sum_{i=1}^{3}|\bl_i-\br|
\end{eqnarray}
We shall be taking ansatz wave functions of the form
\begin{equation} \plabel{psih}
{\mbox{\boldmath $\hat{\eta}$}}_{-}\cdot\br \;\;\Psi_B(\br)\hspace{1cm}
\end{equation}
for $H_1$ hybrid baryons, where $\Psi_B(\br)$ is an exponential function.
It is not difficult to show that ${\mbox{\boldmath $\hat{\eta}$}}_{-}$ lies in the plane spanned
by the three quarks (the ``QQQ plane'').
\section{Quantum numbers of low--lying hybrid baryons}
{\bf Angular Momentum:}
The hamiltonian in Eq. \ref{ham3} is not invariant under rotations in the
junction position $\br$, with fixed quark positions. When the
junction wave function, which is hence not an eigenfunction
of angular momentum,
is combined with the quark motion wave functions,
which are eigenfunctions of angular momentum, it must be done in
such a way that the total angular momentum of the junction and quark motion is
well--defined. Obtaining a well--defined total angular momentum is
a technically challenging problem that is an artifact of
the adiabatic approximation,
which separates junction and quark motion.
We here merely give an intuitive argument why the
total angular momentum $L$ of the $H_1$ baryons is expected to be 1.
The hybrid baryon wave function is proportional to ${\mbox{\boldmath $\hat{\eta}$}}_-\cdot\br$,
and since ${\mbox{\boldmath $\hat{\eta}$}}_-$ lies in the QQQ plane,
it can be regarded as the x--axis, so that
${\mbox{\boldmath $\hat{\eta}$}}_-\cdot\br = \sqrt{\frac{2\pi}{3}} r
(-Y_{11}({\bf \hat{r}}) + Y_{1-1}({\bf \hat{r}}))$
in terms of spherical harmonics. If the mathematics of conservation of
angular momentum is followed through, it is found that if the angular
momentum of the quark motion is $L_q=0$ (corresponding to the lowest energy
quark motion states), then the total angular momentum projection just
equals the angular momentum projection of the junction wave functions,
which in this case is $\pm 1$. Hence the total angular momentum projection
is $\pm 1$ so that $L$ cannot be zero, and should most
likely be 1.
{\bf Exchange symmetry:}
Exchange symmetry transformations $S_{ij}$ exchange
the positions of the quarks $\bl_i \leftrightarrow \bl_j$.
Since the physics does not
depend on the quark position labelling convention, the junction hamiltonian
should be exchange symmetric, as can be seen explicitly in Eq. \ref{ham3},
noting that the junction position $\br$ is
not determined by the positions of the quarks.
We now argue that the junction wave functions of (hybrid) baryons
should transform either totally symmetrically or totally anti--symmetrically
under exchange symmetry. Since the hamiltonian is invariant under
exchange symmetry we have the commutation relation
$[H_{\mbox{\small flux}},S_{ij}]=0$. Combining this with the Schr\"{o}dinger equation
\begin{equation}\plabel{ex} H_{\mbox{\small flux}}\Psi = V(l_1,l_2,l_3)\Psi \hspace{1cm}
\mbox{gives} \hspace{1cm} H_{\mbox{\small flux}} (S_{ij}\Psi) =
V(l_1,l_2,l_3) (S_{ij}\Psi)\end{equation}
so that $S_{ij}\Psi$ is degenerate in energy with $\Psi$.
Now since the baryon and each of the hybrid baryons $H_i$ have
different energies (except when $l_1=l_2=l_3$) it follows
that $S_{ij}\Psi$ must be a multiple of $\Psi$, i.e. that
$S_{ij}\Psi = \varsigma \Psi$, where $\varsigma$ is complex number.
Now note that the product of two exchange symmetry transformations is
the identity, i.e. that
\begin{equation} S_{ij}S_{ij} = 1 \hspace{1cm} \mbox{which implies that}
\hspace{1cm} \varsigma^2 = 1 \end{equation}
or $\varsigma=\pm 1$. Hence $S_{ij}\Psi = \pm \Psi$.
Assume that $S_{12}\Psi = \varsigma\Psi$. We now show\footnote{This result
also follows by noting that $[H_{\mbox{\small flux}},S_{ij}]=0$ implies that
$\Psi$ must be an irreducible representation of the exchange symmetry
group, i.e. totally symmetric, anti--symmetric or mixed symmetry. But since
we already showed that $S_{ij}\Psi = \pm \Psi$, it follows that $\Psi$ is
in either the totally symmetric or anti--symmetric irreducible representation.} that
$S_{23}\Psi =S_{13}\Psi= \varsigma\Psi$, i.e. that $\Psi$
is either totally symmetric or totally anti--symmetric under label
exchange. This follows by the two identities
\begin{eqnarray}
\lefteqn{S_{12}S_{23}S_{13}S_{23}= 1 \hspace{1cm} \mbox{which implies that}
\hspace{1cm} S_{12}\Psi = S_{13}\Psi \nonumber } \\ & & \hspace{-0.7cm}
S_{23}S_{12}S_{13}S_{12}= 1 \hspace{1cm} \mbox{which implies that}
\hspace{1cm} S_{23}\Psi = S_{13}\Psi
\end{eqnarray}
For each of the hybrid baryons $H_i$, there are hence two varieties:
the junction wave function is totally symmetric (S) or totally anti--symmetric (A) under quark label exchange, denoted by
$H_i^S$ and $H_i^A$.
{\bf Parity:}
The inversion of all coordinates $\bl_i\rightarrow -\bl_i$ and
$\br\rightarrow -\br$, called ``parity'', is a symmetry of the junction
hamiltonian in Eq. \ref{ham3}.
${\mbox{\boldmath $\hat{\eta}$}}_{-}$ is a vector in the QQQ plane and is a linear combination
of the $\hat{\bf l}_i$, which span the plane,
with coefficients which are functions of
$l_i$. The lengths
$l_i$ remain invariant under parity.
However, $\hat{\bf l}_i \rightarrow - \hat{\bf l}_i$ under parity.
It follows that
${\mbox{\boldmath $\hat{\eta}$}}_{-}$ is {\it odd under parity}.
The junction wave function
in Eq. \ref{psih} is thus even under parity, since ${\mbox{\boldmath $\hat{\eta}$}}_{-}\rightarrow -{\mbox{\boldmath $\hat{\eta}$}}_{-}$ and $\br\rightarrow -\br$.
For a low--lying hybrid the quark motion wave function is even under
parity, so that the full hybrid baryon wave function has even parity.
Since quarks are fermions, the wave function
should be totally antisymmetric under quark label exchange, called
the Pauli principle.
Since our philosophy is that (hybrid) baryon dynamics is dominated by (non--perturbative) long distance physics, we consider the colour structure of the (hybrid) baryon to be motivated from the long distance limit, i.e. from the strong coupling limit of the hamiltonian formulation of lattice QCD \cite{paton85}. Here, the quarks are sources of triplet colour, which flows along the string connected to the quarks into the junction, where an $\epsilon_{ijk}$ neutralizes the colour. The colour wave function $\epsilon_{ijk}$ is hence totally antisymmetric under exchange of quarks for {\it both} the conventional and hybrid baryon.
This imposes constraints on the combination of
flavour and non--relativistic spin $S$ of the three quarks that is allowed.
For a totally symmetric hybrid baryon
junction wave function, the flavour--spin wave
functions must be totally symmetric. This is because we are interested
in the low--lying hybrid baryons which have the
quark motion wave function in ground state, i.e. totally symmetric. If the
flavour is $\Delta$, which is totally symmetric, this implies that the spin
must be totally symmetric, i.e. $S=\frac{3}{2}$. Similarly for flavour
$N$ the spin must be $\frac{1}{2}$. For a totally antisymmetric junction
wave function, the flavour--spin wave function must be totally antisymmetric.
For $\Delta$ flavour this implies that the spin must be totally antisymmetric,
which is not realizable. Hence there are no $\Delta$ hybrid baryons with
totally antisymmetric junction wave functions. The $N$ flavour is found to have
spin $\frac{1}{2}$.
The quantum numbers of the lowest--lying states that can be constructed
on the $H_1$ adiabatic surface are indicated in Table \ref{tabqu}.
The total angular momentum ${\bf J} = {\bf L} + {\bf S}$. Since $L=1$ for
ground state $H_1$ hybrid baryon, $J=\frac{1}{2},\frac{3}{2}$ for $S=\frac{1}{2}$,
and $J=\frac{1}{2},\frac{3}{2},\frac{5}{2}$ for $S=\frac{3}{2}$.
These assignments are indicated in Table \ref{tabqu}.
One notes from Table \ref{tabqu} that amongst the $H_1^S$ hybrid baryons,
there are $N \frac{1}{2}^+$ and $\Delta \frac{3}{2}^+$
states which have identical quantum numbers to the conventional $N$ and
$\Delta$ baryons.
\begin{table}[t]
\begin{center}
\caption{\small Quantum numbers of low--lying hybrid baryons for the adiabatic surface $H_1$. In the absense of spin dependent forces all these states are degenerate. $N,\Delta$ are the flavour structure of the wave function (i.e. those of the conventional baryons $N,\Delta$ respectively) and $P$ the parity. }
\label{tabqu}
\begin{tabular}{|c||l|c|l|}
\hline
Hybrid Baryon & $L$ & $S$ & $(N,\Delta)^{2S+1}J^P$ \\
\hline
$H_1^S $ & 1 & $\frac{1}{2},\frac{3}{2}$ & $N^2 {\frac{1}{2}}^+, \; N^2 {\frac{3}{2}}^+, \; \Delta^4 {\frac{1}{2}}^+, \; \Delta^4 {\frac{3}{2}}^+, \; \Delta^4 {\frac{5}{2}}^+$\\
$H_1^A $ & 1 & $\frac{1}{2}$ & $N^2 {\frac{1}{2}}^+, \; N^2 {\frac{3}{2}}^+$\\
\hline
\end{tabular}
\end{center}
\end{table}
It is interesting to compare our hybrid baryons to the predictions
of the bag model. Out of all the states listed under
$H_1^S$ and $H_1^A$ in Table \ref{tabqu},
only one pair of
$N^2 {\frac{1}{2}}^+, \; N^2 {\frac{3}{2}}^+$ states have the same
flavour, spin $S$, total angular momentum and parity as
the low--lying hybrid baryons in the bag model \cite{bag}.
In fact, for the $H_1^S$
hybrid baryons, the bag model swaps the $N$ and $\Delta$ flavours
from our assignments, keeping other quantum numbers the same.
Both our model and the bag model has seven low--lying hybrid baryons
\cite{bag}.
\section{Numerical estimate of the hybrid baryon mass}
The difference between the hybrid and conventional baryon adiabatic potentials
(or junction energies)
as a function of quark positions, $V_{H_1}(l_1,l_2,l_3)-V_B(l_1,l_2,l_3)$,
was determined numerically
from the first part of Eq. \ref{ex} by using the
hamiltonian in Eq. \ref{ham3}, and were displayed in ref. \cite{conf}.
Now define the
hybrid baryon potential as
\begin{equation}\plabel{pot2}
\mbox{\={V}}_{H_1} (l_1,l_2,l_3) \equiv
\mbox{\={V}}_B (l_1,l_2,l_3) + V_{H_1} (l_1,l_2,l_3) - V_B (l_1,l_2,l_3)
\end{equation}
where $\mbox{\={V}}_B (l_1,l_2,l_3)$ is the phenomenologically
successful relativized baryon hamiltonian with Coulomb and
linear potential terms of ref. \cite{capstick86} (with spin--spin, spin--orbit
and tensor interactions neglected); and the parameters are
also those of ref. \cite{capstick86}.
Note that the Coulomb interaction of the conventional and
hybrid baryon is assumed to be identical.
We solve the Schr\"{o}dinger equation for the hamiltonian in Eq.
\ref{pot2} with 95 spin--space basis states incorporating $L_q =0,1,2$
harmonic oscillator wave functions for the $J=\frac{1}{2}$ case,
i.e. construct 95 $\times$ 95 dimensional matrices.
These matrices are subsequently diagonalized.
The differences between the energies for the hybrid and the conventional baryon is then
added to the experimental mass of the lowest baryon,
taken as the spin--averaged mass of the $N$ and $\Delta$, i.e. 1085 MeV \protect\cite{pdg98}. The first three quark orbital
excitations $L_q=0,1,2$ of hybrid baryons composed of up and down quarks
are found to have masses 1976, 2341 and 2619 MeV respectively.
Hence, for the lowest hybrid baryon level, with the quantum numbers in
Table \ref{tabqu},
we obtain that $M_{H_1} - M_B = 891$ MeV, giving a mass estimate of
$M_{H_1} = 1976$ MeV.
This mass estimate is substantially higher than other mass estimates in the literature:
$\sim 1.5$ GeV in the bag model \cite{bag} and $1.5\pm 10\%$ GeV in QCD sum rules \cite{zpli}.
There are two crucial assumptions that were made in the early work on (hybrid) meson masses in
the flux--tube model: the adiabatic motion of quarks and the small oscillation approximation for
flux motion \cite{paton85}. It was later shown that when the adiabatic approximation is lifted,
the masses goes up, and when the small oscillation approximation is lifted, the masses goes down
\cite{paton85}. In our study of (hybrid) baryons we have partially lifted the adiabatic approximation
by working in the centre of mass frame. We have fully lifted the small oscillation approximation.
The effects on the masses of (hybrid) baryons when the various approximations are lifted are the same
as those found for (hybrid) mesons.
In our simulation, we obtain the average values $\sqrt{\langle\rho^2\rangle}=\sqrt{\langle\lambda^2\rangle} = 2.12,\; 2.52$ GeV$^{-1}$
for the low--lying baryon and $H_1$ hybrid baryon respectively.
${\langle\rho^2\rangle}={\langle\lambda^2\rangle}$ is expected since the spatial parts of the
wave functions of the low--lying states
are totally symmetric under exchange symmetry. The hybrid baryon is 20\% larger than the
conventional baryon.
\section{Phenomenology}
The sign of the the Coulomb interaction
is expected to be the same for both conventional and hybrid baryons
\cite{bag}.
This means that
the hyperfine interaction has the same sign in both situations, so that
the $\Delta$ hybrid baryons are always heavier than the $N$ hybrids.
This implies that only four of the original seven low--lying baryons,
the $N$ hybrids, are truely low--lying.
We expect {\it a priori} the most phenomenologically interesting decay of the low--lying hybrid baryons to be the P--wave decay to $N\rho$ and $N\omega$, simply because the phase space is favourable and $\rho$ and $\omega$ are easily isolated experimentally. The $N\rho$ decay would be especially relevant to the electro-- and photoproduction of hybrid baryons at TJNAF via the vector meson dominated coupling of the photon to the $\rho$. Indeed, a search for excited $N^*$ resonances with mass $< 2.2$ GeV is currently underway in Hall B \cite{kees}. Given the mass estimate for the low--lying hybrid baryons, the detection of hybrid baryons in $N\rho$ or $N\omega$ is feasible at TJNAF.
There are also planned experiments in $\pi N$
scattering by Crystal Ball E913 at the new D--line at Brookhaven
with the capability of searching for states
in $N\{\eta,\rho,\omega\}$, which would isolate states in the mass region
$\sim 2$ GeV \cite{bris}.
The decay $\psi\rightarrow p\bar{p}\omega$ has been observed with a branching ratio of $1.30\pm 0.25\; 10^{-3}$ and $\psi\rightarrow p\bar{p}\eta^{'}$ with branching ratio $9\pm 4\; 10^{-4}$ \cite{pdg98}. Since gluonic hadron production is expected to be enhanced above conventional hadron production in the glue--rich decay of the $\psi$, it is possible that a partial wave analysis of the $p\omega$ or $p\eta^{'}$ invariant masses would yield evidence for hybrid baryons. Future work at BEPC and an upgraded $\tau$--charm factory would be critical here.
\section{Conclusions}
The spin and flavour structure of the low--lying
hybrid baryons have been specified,
and differ from their structure in the bag model.
Exchange symmetry constrains the spin and flavour of the (hybrid) baryon
wave function. The orbital angular momentum of the low--lying hybrid baryon
is argued to be unity, with the parity even, contrary to conventional
baryons where $L=1$ would imply the parity to be odd.
The low--lying hybrid baryon adiabatic potential and
mass has been estimated numerically. The mass estimate is considerably
higher than bag model and QCD sum rule estimates.
|
\section{Introduction}
\label{sec:introduction}
\indent\indent In this paper, we consider extensions of the bar-attendance problem introduced by Arthur \cite{arthur94} and simplified into a minority game by Challet and Zhang \cite{challet9798}. In its simplest form, the minority game mimics the internal dynamic of the exchange of one commodity. Agents are allowed to buy or sell this commodity at each time step. No attempt is made to model any external factors that influence the market. Here, we introduce symmetric and asymmetric three sided games as extensions of the minority game.
In the symmetric three sided model, the agents have to choose between three identical sides at each time step. These three sides are trading with each other, agents on one side buying from the second side to sell to the third. This model mimics the cyclic trading of goods. If we group any two sides together and consider the trading between this imaginary group and the third side, the model reduces to a kind of minority game with an uneven distribution of the agents. Hence, the connection between this model and the minority game is very strong.
In the asymmetric three sided model, the agents can buy or sell a commodity at each time step, but they can also be inactive, that is, they are allowed to miss a turn. In contrast to the symmetric model, the three choices are not equivalent, as being inactive appears as a compromise between buying and selling. This model can be thought of as an open minority game in the sense that the agents buying and selling are playing a minority game with a variable number of agents at each turn.
In Sec. \ref{sec:themodels}, the minority game is briefly recalled and the two new three sided models are described in detail. In Sec. \ref{sec:symmetric}, the symmetric three sided model is numerically investigated, while in Sec. \ref{sec:asymmetric}, the asymmetric three sided model is investigated. Sec. \ref{sec:conclusions} presents a comparison between the minority game and the two three sided models, as well as our conclusions.
\section{The models}
\label{sec:themodels}
\indent\indent In the minority game, an odd number $N$ of agents have to choose between two sides, $1$ or $2$, at each time step. An agent wins if he chooses the minority side. The record of which side was the winning side for the last $m$ time steps constitutes the history of the system. The agents analyze the history of the system in order to make their next decision.
In the symmetric three sided model, a number of agents $N$ have to choose between three sides, $1$, $2$ or $3$, at each time step. $N$ is not a multiple of 3. The agents on side 1 buy from side 2 to sell to side 3, the agents on side 2 buy from side 3 to sell to side 1 and the agents on side 3 buy from side 1 to sell to side 2. This cyclic trading pattern is shown in Fig. 1. It is assumed that the profit or loss at a side is reflected in the difference between the number of agents they are selling to and the number of agents they are buying from. For instance, $N_3 -N_2$ is a measure of the profit of side 1. Agents choosing the side with the highest profit win and are rewarded with a point. Agents choosing the side with the lowest profit lose and consequently lose a point. Agents choosing the side with the intermediate profit neither lose nor gain a point. Agents strive to maximize their total number of points.
\end{multicols}
\begin{figure}[h]
\centering\begin{picture}(5,5)(-2,1)
\put(-0.2,4.9){$N_1$}
\put(0,5){\circle{2}}
\put(-1.932,1.9){$N_2$}
\put(-1.732,2){\circle{2}}
\put(1.532,1.9){$N_3$}
\put(1.732,2){\circle{2}}
\put(0.732,2){\vector(-1,0){1.364}}
\put(-1.232,2.866){\vector(1,2){0.682}}
\put(0.5,4.134){\vector(1,-2){0.682}}
\put(-3.5,5){buying from 2}
\put(-4,3){selling to 1}
\put(-3,0.8){buying from 3}
\put(1,0.8){selling to 2}
\put(2,3){buying from 1}
\put(1,5){selling to 3}
\end{picture}
\caption{Schematic representation of the symmetric three sided model. The arrows indicate the direction of the exchange. $N_i$ is the number of agents choosing side $i$.
\label{fig1}}
\end{figure}
\begin{multicols}{2}
In the asymmetric three sided model, a number of agents $N$ also have to choose between three sides, $1$, $2$ or $3$, at each time step. $1$ corresponds to selling, $2$ to doing nothing and $3$ to buying. The agents buying or selling are said to be active, while the agents doing nothing are said to be inactive. The agents choosing the smaller group among buyers and sellers win and are rewarded with a point. The agents choosing the larger group among buyers and sellers lose and they lose a point. The points of the inactive agents don't change. If there is the same number of buyers and sellers, the points of all the agents remain unchanged. However, the inactive agents are recorded as winners in the history of the system, on the grounds that they achieved the same result as the buyers and sellers, but without taking any risk. Again, agents strive to maximize their total number of points.
\end{multicols}
\begin{figure}
\begin{minipage}[b]{.46\linewidth}
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
History & $\sigma$ & $\sigma'$\\
\hline
(1,1) & 3 & 2 \\
(1,2) & 1 & 3 \\
(1,3) & 2 & 2 \\
(2,1) & 2 & 1 \\
(2,2) & 3 & 3 \\
(2,3) & 1 & 1 \\
(3,1) & 3 & 1 \\
(3,2) & 2 & 2 \\
(3,3) & 3 & 1 \\
\hline
\end{tabular}
\end{center}
\vfill
\end{minipage}
\hfill
\begin{minipage}[b]{.46\linewidth}
\noindent Table 1: The first column lists all the possible histories of the system for the last 2 time steps $(m=2)$. A strategy is a set of decisions for all the different possible histories. Two example strategies $\sigma$ and $\sigma'$ are shown in the second and third columns.
\end{minipage}
\end{figure}
\begin{multicols}{2}
In each model, the record of which side was the winning side for the last $m$ time steps constitutes the history of the system. For a given $m$, there are $3^m$ different histories. The 9 different histories for $m=2$ are listed in the first column of table 1. Every agent makes a decision for the next time step according to the history of the system. To be able to play, an agent must have a strategy that allows him to make a decision for any of the $3^m$ different histories. The second and third columns of table 1 list two possible sets of decisions, $\sigma$ and $\sigma'$, that we will call strategies.
Each agent has at his disposal a fixed set of $s$ strategies chosen at random, multiple choices of the same strategy being allowed. At any one moment in time, the agent only uses one of these strategies to make a decision. To allow an agent to decide which strategy to use, every strategy is awarded points, which are called virtual points. The virtual points of a strategy are the points the agent thinks he could have earned had he played with this strategy. Hence, the virtual points are rewarded using the same scheme as the points given to the agents, the prediction of a strategy being compared to the actual decisions. A strategy predicting the winning side is awarded a virtual point, a strategy predicting the losing side loses a virtual point and a strategy predicting the third side does not gain or lose any points. In the asymmetric model, in the case of an equal number of buyers and sellers, the virtual points of all strategies remain unchanged. An agent always plays with the strategy with the highest number of virtual points. When more than one strategy has the highest number of virtual points, one of them is chosen at random.
If we compare two strategies $\sigma$ and $\sigma'$ component by component, we see that for some histories they can make the same prediction and for others they can make different predictions. In the example in table 1, the decisions differ when the history is (1,1), (1,2), (2,1), (3,1) and (3,3). To consider this feature, we have to distinguish between the symmetric model and the asymmetric one. For the former, the three sides are equivalent and only the number of differences between the strategies can give a measure of the difference between two strategies in the strategy space. For the latter, there is a qualitative difference between the three sides. This qualitative difference should appear in the definition of the difference between strategies.
Consider first the symmetric three sided model. As the three sides are equivalent, a geometrical representation should put them at the same distance from one another. A convenient measure of the differences between two strategies $\sigma$ and $\sigma'$ is
\begin{equation}
d_s = {1\over 3^m} \sum_{i=1}^{3^m} \delta (\sigma_i - \sigma'_i )
\label{eq:symmetric distance definition}
\end{equation}
where $\delta (0) = 1$, and $\delta (x) = 0$ otherwise. $d_s$ is defined as the distance between strategies in the symmetric model. This definition takes into account the geometrical structure of the strategy space, including the equivalence between the three sides. In the example of table 1, $d_s =5/9$. By definition, the symmetric distance is a number ranging from 0 to 1.
As Eq. (\ref{eq:symmetric distance definition}) shows, the symmetric distance $d_s$ is defined as a sum of $3^m$ terms, which we label $d_s^{(i)}$'s. Each of these terms is equal to 0 with probability 1/3 or equal to 1 with probability 2/3. The average distance between two strategies is $\overline{d}_s = 2/3$, while the variance of the symmetric distance distribution is $\sigma^2_s = 2/3^{m+2}$. The symmetric distance between two strategies corresponds to the probability that these two strategies will give different predictions, assuming that all the histories are equally likely to occur. The symmetric distance corresponds to the distance defined in the minority game \cite{dhulst99}. Two strategies at $d_s = 0$ are correlated, two strategies at $d_s =2/3$ are uncorrelated and two strategies at $d_s = 1$ are anticorrelated.
In the asymmetric three sided minority game, selling is just the opposite decision to buying while doing nothing is a compromise. Consequently, the normalized asymmetric distance between strategies $\sigma$ and $\sigma'$,
\begin{equation}
d_a = {1\over 3^m} \sum_{i=1}^{3^m} {| \sigma_i - \sigma'_i |\over 2}
\label{eq:asymmetric distance definition}
\end{equation}
is a measure of the difference between the two strategies. $d_a$ is defined as the distance between strategies in the asymmetric model. This definition takes into account the fact that buying is more different from selling than it is from being inactive in an arbitrary way. In the example of table 1, $d_a =4/9$. By definition, the asymmetric distance is a number ranging from 0 to 1.
As shown by Eq. (\ref{eq:asymmetric distance definition}), the asymmetric distance $d_a$ is defined as a sum of $3^m$ terms we label $d_a^{(i)}$'s. When the component of a strategy is equal to 2, this component can never give a $d_a^{(i)}$ equal to 1. In other words, the inactive side has no side at distance 1 from itself. Considering all the possibilities, the probability to find a $d_a^{(i)}$ of 0 is 1/3, of 0.5 is 4/9 and of 1 is 2/9. The average asymmetric distance between strategies is $\overline{d}_a = 4/9$, while the variance of the asymmetric distance distribution is equal to $\sigma^2_a = 11/3^{m+4}$. The interpretation of this asymmetric distance is ambiguous. In fact, the opposite to selling is buying, but the opposite to being inactive is being inactive. Hence, $d_a$ is not a measure of the probability that two strategies would give opposite decisions. Two correlated strategies are at $d_s = 0$ from each other, two uncorrelated strategies are at $d_a = 1/2$ from each other, but two anticorrelated strategies can be at $d_a = 0$ or $d_a = 1$ from each other.
\section{Numerical results for the symmetric model}
\label{sec:symmetric}
\indent\indent In this section, we report on numerical investigations of the properties of the symmetric three sided model, interpreting the results using the symmetric distance.
Fig. 2 presents a typical result for the time evolution of the attendance at one side. The simulation is for $N=101$ agents with $s=2$ strategies each and a memory of $m=3$. The result for the attendance at one side is very similar to the results of the minority game, the mean attendance being shifted to $N/3$ instead of $N/2$. Given an agent choosing one side, the average distance between the strategy used by this agent and the strategies used by the other agents is $\overline{d} = 2/3$. That is, around $2/3$ of the agents should choose one of the two other sides. Hence, the average attendance at one side is $N/3$.
\end{multicols}
\begin{figure}[h]
\centerline{\epsfig{file=fig2.eps,
width=15cm,height=8cm}}
\caption{Numerical simulation of the attendance at one side for the symmetric three sided model. The choice for the parameters is $N=101$, $s=2$ and $m=3$.}
\label{fig2}
\end{figure}
\begin{multicols}{2}
The variance of the attendance at one side as a function of the size of the memory $m$ is presented at Fig. 3 for $N=101$ agents with $s=2$ strategies. The result is again very similar to the minority game, with a very high variance for $m<3$, a minimum at $m=3$ and the variance going to $2N/9$, the random value, as $m$ goes to infinity. Curves of the same shape are obtained for the variance of the number of winners or the variance of the number of losers. Also, the maximum profit or the number of agents on the more crowded side exhibit the same behaviour. Each of these curves has a minimum for $m=3$. For $m<3$, the number of strategies used at each time step is a representative sample of the space of the strategies. Consequently, the variance of the attendance is directly related to the variance of the distance distribution, $\sigma^2_s = 2/3^{m+2}$. In fact, the variance of the attendance scales like $1/ \sigma^2_s$. On the contrary, for $m>3$, the space of the strategies is very large, so that most of the strategies used are uncorrelated. As a result, the kinetics of the system are the same as the kinetics of a random walk. Between these two behaviours, for $m$ around 3, the agents organize themselves better, a crowd-anticrowd effect being obtained \cite{johnson98-1}.
\end{multicols}
\begin{figure}[h]
\centerline{\epsfig{file=fig3.eps,
width=15cm,height=8cm}}
\caption{Variance of the number of agents at one side as a function of the size of the memory $m$ ($\times$) for $N=101$ and $s=2$ (symmetric model). The analytical results using the calculation of Johnson {\it et al.} [4] is also shown ($\Box$).}
\label{fig3}
\end{figure}
\begin{multicols}{2}
Even for small values of $m$, the space of strategies is very large, of size $3^{3^m}$. But as in the minority game, not all the strategies are uncorrelated. If we suppose that $1/ \sigma^2_s$ gives an estimate of the number of uncorrelated strategies, the method of Johnson {\it et al.} \cite{johnson98-1} can be used to find an analytical expression for the variance of the attendance at one side. We followed the original calculation in \cite{johnson98-1}, with a size $a = 3^{m+2}/2$ for the space of strategies and a variance of 2/9 for an independent agent. The analytical result obtained by this method is compared in Fig. 3 to the result of the numerical simulations. The curves agree qualitatively.
\end{multicols}
\begin{figure}
\begin{tabular}{cc}
\begin{picture}(6,2)(0,0)
\epsfig{figure=fig4a.eps,width=0.5\linewidth}
\end{picture}
&
\begin{picture}(8,2)(0,0)
\epsfig{figure=fig4b.eps,width=0.5\linewidth}
\end{picture}
\end{tabular}
\caption{(a) Average points earned by the agents and their strategies as a function of the time for $N=101$, $s=2$ and $m=3$. The ordinate on the left refers to the points of the agents (bold line) while the ordinate on the right refers to the points of their strategies (simple line). (b) Profit rate of the strategies as a function of $m$ for $N=101$ and $s=2$.}
\label{fig4}
\end{figure}
\begin{multicols}{2}
Fig. 4 (a) presents a typical result for the average number of points given to the agents and their strategies. The parameters of the simulation are $N=101$, $s=2$ and $m=3$. Note that there are two different ordinate scales. As Fig. 4 (a) shows, the virtual points are steadily decreasing with time. In contrast, the points given to the agents display a more complex behaviour. The points given to the agents increase very slowly for $m<3$ and then oscillate around 0 for $m>3$. There seem to be no special behaviour for $m=3$. The time evolution of the virtual points can be approximated by a linear relation with a profit rate $\tau$. We define $\tau$ as the average number of points earned by a strategy at each time step. Fig. 4 (b) presents $\tau$ as a function of the memory $m$ for $N=101$ and $s=2$. For $m<3$, the strategies are slowly losing points, the worst results being obtained for $m=3$. For $m>3$, the virtual points oscillate around 0. Hence, the agents seem to be able to choose their strategy efficiently, in the sense that the strategies they choose win more often than the average strategy. This behaviour is to be contrasted with the minority game where the agents are not able to choose a strategy efficiently.
As a summary, the symmetric model is a direct extension of the minority game to three sides. The results found are very similar, with a glassy phase transition \cite{savit97} when the size of the memory of the agents is increased. We numerically identified a critical value $m_c$ for the size of the memory. For $m<m_c$, the space of strategies is crowded and its geometrical structure is apparent in the results. As this structure is encoded in the distance definition, the system is driven by its distance distribution . For $m>m_c$, the number of strategies used is not relevant as most of the strategies used are uncorrelated. The kinetics of the system reduce to agents choosing one of the three sides at random. Hence, there is a transition from a system driven by its distance distribution to a random system.
\section{Numerical results for the asymmetric model}
\label{sec:asymmetric}
\indent\indent We investigated numerically the different properties of the asymmetric three sided game. In the figures, 1 denotes buying, 2, doing nothing and 3, selling.
The attendance of the three different sides as a function of the size of the memory $m$ is plotted at Fig. 5 for $N=101$ agents, playing with $s=2$ strategies each. The number of agents in the winning side is also presented. For small $m$ values, most of the agents are buying or selling (the two superimposed upper curves). Just a few of them are doing nothing (the lower curve for small $m$ values). As the size of the memory is increased, the system corresponds more and more to the agents guessing at random between the three possibilities. Also, for small values of $m$, the number of winners is significantly more than 1/3, the random guess value. Fig. 5 is interesting because the difference between the three sides is clearly apparent.
\end{multicols}
\begin{figure}[h]
\centerline{\epsfig{file=fig5.eps,
width=15cm,height=8cm}}
\caption{Numerical result for the average attendances at the three sides as functions of $m$ for $N=101$ and $s=2$. The two upper curves ($\triangle$ and $\bigcirc$) are for the selling and buying options. The lower curve ($\times$) is for the inactive agents. The average number of winners is also presented ($\Box$) as a function of $m$ for the same choice of parameters.}
\label{fig5}
\end{figure}
\begin{multicols}{2}
In Fig. 6, the variance of the attendances at the three sides and the variance of the number of winners are presented as functions of $m$ for $N=101$ and $s=2$. For $m<6$, the variance of the number of inactive agents is significantly higher than $2N/9$, the value for agents guessing at random. The variances of the number of buyers and sellers has a minimum at $m=2$. The variances of the three sides increase to $2N/9$ as $m$ increases. Hence, there seems to be an organization of the agents for $m$ around 2. The variance of the number of winners has a shape very similar to the one found in the minority game. For small value of $m$, the variance diverges like a power law of $m$; at $m\simeq 7$, it seems to reach a minimum and for higher values of $m$, it goes asymptotically to a value near $N/9$. However, the existence of a minimum at $m=7$ could not be confirmed unequivocally by the numerical simulations.
\end{multicols}
\begin{figure}[h]
\centerline{\epsfig{file=fig6.eps,
width=15cm,height=8cm}}
\caption{Variances of attendances at the three sides as functions of $m$ for $N=101$ and $s=2$. $\triangle$ and $\bigcirc$, selling and buying options, $\times$ inactive agents. $\Box$ is the variance of the number of winners as a function of $m$ for the same choice of parameters.}
\label{fig6}
\end{figure}
\begin{multicols}{2}
Fig. 5 and 6 show that for small $m$ values, the behaviour of the system is directly related to the properties of the distance distribution. The proportion of people buying or selling is of the same order as the average distance, 4/9, while the variance of the number of winners scales as $1 /\sigma^2_a$, the inverse of the variance of the asymmetric distance distribution. These properties were also present in the minority game. In this asymmetric model, the variance of the attendance at one side does not represent the wasted number of points. The wasted number of points is defined to be the difference between the maximum points that can be earned by the system at each time step and the average points actually earned by the system at each time step. This is why we also have to consider the properties of the number of winners in addition to the properties of the attendances.
For higher values of $m$, the strategy space is so large that most of the used strategies are uncorrelated. The system is similar to a system with agents choosing at random from the three sides. In the minority game, the relative attendance predicted by the distance distribution is the same as the one predicted by random guesses, that is 1/2. In the present three sided minority game, these two ratios are 4/9 and 1/3 respectively. Hence, the transition from a system driven by the distance distribution to a system of agent guessing at random is seen directly in the attendance of the different sides.
Fig. 7 presents the average success rate of one side, that is, the probability that at any one moment in time, one side will win. As expected, the sides corresponding to buying and selling are symmetric and more likely to win than the inactive side. In fact, there are $(N+1)(N+2)/2$ different configurations for the attendances of the 3 sides. Among these, only $(N+2)/2$ make the inactive agents winners if $N$ is even, $(N+1)/2$ if $N$ is odd. Hence, if all the situations were equally likely to occur, the inactive agents would win at most about every $N+1$ time steps. This is the order of the asymptotical value for the success rate of this side. For low values of $m$, the success rate of the inactive side is higher than the asymptotical value, implying that the agents playing are organizing themselves rather well. The transition between organized and non-organized agents is for $m=2$ in Fig. 7.
\end{multicols}
\begin{figure}[h]
\centerline{\epsfig{file=fig7.eps,
width=15cm,height=8cm}}
\caption{Success rates of the three different sides as functions of $m$ for $N=101$ and $s=2$. The two upper curves ($\triangle$ and $\bigcirc$) are for selling and buying while the lower curve ($\times$) is for the inactive agents.}
\label{fig7}
\end{figure}
\begin{multicols}{2}
Fig. 8 confirms the organization of the agents. The profit rates of the agents and their strategies are shown as functions of $m$. We define a profit rate as the average number of points earned at each time step. For values of $m$ less than $m=5$, the agents are able to choose strategies which are more successful than the average ones. On the contrary, for $m>5$, they are doing worse than guessing at random. The curve of the profit rate of the strategies suggests that the transition takes place for $m=2$.
\end{multicols}
\begin{figure}[h]
\centerline{\epsfig{file=fig8.eps,
width=15cm,height=8cm}}
\caption{Average profit rates of the agents and their strategies as functions of $m$ for $N=101$ and $s=2$. $\Box$ is for the profit rate of the agents while $\bigcirc$ is for the profit rate of their strategies.}
\label{fig8}
\end{figure}
\begin{multicols}{2}
As a summary, in the asymmetric three sided minority game, agents playing with a small memory win more points on average than agents playing with a bigger memory in a pure population, that is, a population of all the agents with the same memory size $m$. As in the minority game and the symmetric three sided model, a glassy phase transition \cite{savit97} is found at a particular value of the memory $m_c$. For $m<m_c$, the geometrical properties of the space of strategies is apparent, especially the asymmetry between the three sides. Most of the agents are playing and the system is driven by the distance distribution. In contrast to the minority game, this property is seen directly in the number of agents on each side. For $m>m_c$, the strategies used are uncorrelated and the system is similar to a system of agents guessing at random. Considering the adaptation of the agents, they are unable to realize that the wiser choice is to decide to be inactive. In fact, more than half of the active agents will lose. The agents are fooled because they base their confidence in virtual points, not on their profit. Hence, the agents are always tempted to play even if they are unlikely to win.
\section{Conclusions}
\label{sec:conclusions}
\indent\indent We introduced two three sided models as extensions of the minority game. In the symmetric three sided model, agents are given three equivalent choices while in the asymmetric three sided model, agents have the opportunity to miss a turn and not play. We have investigated these two new models numerically and compared the results with the original minority game.
In both models, we defined a distance between the strategies of the agents. These distances incorporate in their definitions the geometrical structure of the space of strategies. In the symmetric model, the geometrical structure of the space of strategies is very similar to the one in the minority game. The distance gives a measure of the correlation between two strategies. Conversely, the distance in the asymmetric model has no obvious interpretation.
A transition between a system driven by the distance distribution and a system of agents guessing at random was identified numerically in both models. However, in contrast to the minority game, the agents make their highest profit for $m$ small and not at the transition value of $m$. In the distance driven phase, the agents organize themselves, as in the minority game. In contrast to the minority game, however, the average profit rate of the agents is higher than the average profit rate of the strategies, indicating that the agents are choosing their strategies efficiently. In the symmetric model, the transition is apparent in the variance of the number of agents choosing one side while in the asymmetric model the transition is seen in the number of agents itself. This latter property of the asymmetric model is a direct consequence of the geometrical structure of the space of strategies.
In the future, we intend to investigate both models analytically. The symmetric model, in particular, should be amenable to analytical treatment, perhaps following the methods introduced in \cite{challet99} for the two sided minority game.
|
\section{Introduction}
$B$ meson decays with non-charmed final states will play an important role
in the detailed investigations at future $B$--factories. Among these
processes the ones involving the quark transitions
$b \to u \ell \bar{\nu}_\ell$ and $b \to s \gamma$ are of interest
with respect to the determination of $V_{ub}$ and $V_{ts}$ as well
as to discover or constrain effects of physics beyond the Standard
Model (SM).
From the theoretical side a lot of progress has been made by employing
an expansion in inverse powers of the heavy quark mass $m_b$. Using
operator product expansion (OPE) and the symmetries of Heavy Quark
Effective Theory (HQET) \cite{HQET} the nonperturbative
uncertainties can be reduced
to a large extent in inclusive decays \cite{inclusive}. Concerning
the transitions mentioned above the inclusive semileptonic or radiative
processes such as $B \to X_u \ell \bar{\nu}_\ell$ and $B \to X_s \gamma$
are the ones which we shall address in the present paper.
Looking at decay spectra of the leptons and photons in these processes,
it has been pointed out that
in certain regions of phase space (such as the endpoint region of the
lepton energy spectrum in $B \to X_u \ell \bar{\nu}_\ell$ or the photon
energy spectrum in $B \to X_s \gamma$ close to the endpoint of
maximal energy) the correct description requires more than
the naive $1/m_b$ expansion \cite{Neubert,Bigietal,MannelNeubert}.
In these endpoint regions
it is required to sum the leading twist
contributions into a non-perturbative ``shape'' function, which describes the
distribution of the light-cone component of the $b$ quark residual
momentum inside the $B$ meson.
Unfortunately, there is no way to avoid these endpoint regions due to
experimental cuts. In $B \to X_u \ell \bar{\nu}_\ell$ cuts on the lepton
energy and/or on the hadronic invariant mass of the final state are
required to suppress the much larger charm contribution
\cite{BaBarBook,Uraltsev,Falk}. Likewise, in
$B \to X_s \gamma$ a cut on the photon energy is mandatory to reduce the
background from ordinary bremsstrahlung in inclusive $B$ decays
\cite{bsgammaexp}.
These cuts more or less reduce the accessible part of phase space to the
endpoint regions. Hence it is unavoidable to get a theoretical handle
on this light-cone distribution function.
The distribution function is universal, since it depends only on the
properties of the initial-state $B$ hadron. This fact may be used to
establish a model independent relation between the two inclusive decays
$B \to X_u \ell \bar{\nu}_\ell$ and $B \to X_s \gamma$ \cite{Neubert1}.
Parametrizations of this function have been proposed, which include
the known features such as the few lowest moments
\cite{BaBarBook,MannelNeubert}. It has also been
shown that the popular ACCMM model \cite{ACCMM}
for a certain range of its parameters
is indeed consistent with this QCD based approach \cite{BigiUraltsev}.
There have been various suggestions how to overcome the non-perturbative
uncertainties induced by the distribution function. One way is to
consider moments of appropriate distributions
\cite{Neubert,Bigietal,MannelNeubert,Bauer,WiseLigeti}.
The first few moments
are then sensitive only to the first few moments of the distribution
function which are known. However, due to experimental cuts only parts
of the distributions can be measured such that only moments involving
cuts can be obtained. Depending on the cut,
these quantities are sensitive to large
moments of the distribution function and in this way the non-perturbative
uncertainties reappear.
In the present paper we propose a different approach. We suggest to directly
compare the light-cone spectra of the final state hadrons in
$B \to X_u \ell \bar{\nu}_\ell$ and $B \to X_s \gamma$. The individual
rates depend on the shape function while the ratio of the two spectra
is not very sensitive to non-perturbative effects
even if cuts are included.
We discuss the kinematic variables which allow a direct
comparison of the two processes and consider the radiative
corrections entering their relation.
In the next section we give a
derivation of the non-perturbative light cone distribution. Based on this
we calculate in section 3 the perturbative corrections and combine these
with the non-perturbative light cone distribution function. Finally we
study the comparison between $B \to X_u \ell \bar{\nu}_\ell$ and
$B \to X_s \gamma$ quantitatively and conclude.
\section{Non-perturbative Contributions:
Light cone distribution function of the heavy quark}
Although the derivation of how the light-cone distribution function
emerges in the heavy-to-light transitions at hand is well known
\cite{Neubert,MannelNeubert,Bigietal}, we consider it useful to rederive it
here, since our discussion of radiative corrections will be based on
this derivation.
We are going to consider heavy to light transitions such as $b \to u$
and $b \to s$ decays. The relevant effective Hamiltonian is obtained from
the Standard Model by integrating out the top quark and the weak bosons.
The dominant QCD corrections are taken into account as usual by
running this effective interaction down to the scale of the bottom
quark. The corresponding expressions are known to next-to-leading order
accuracy \cite{Buras,AliGreub,ChetMM} and one obtains for the relevant pieces
\begin{eqnarray}
&& H_{eff}^{\rm sl} = \frac{G_F}{\sqrt{2}}
(\bar{b} \gamma_\mu (1-\gamma_5) u)(\bar{\nu} \gamma_\mu (1-\gamma_5) \ell)
\\
&& H_{eff}^{rare} = \frac{G_F}{\sqrt{2}} \frac{e}{8 \pi^2} C_7 (\mu)
m_b (\mu) (\bar{b} \sigma_{\mu \nu} (1-\gamma_5) s)|_\mu F^{\mu \nu}
\end{eqnarray}
mediating semileptonic $b \to u$ transitions and
radiative $b \to s$ decays respectively. $C_7 (\mu)$ is a Wilson
coefficient, depending on the renormalization scale $\mu$,
and $F^{\mu \nu}$ is the electromagnetic field strength.
In the following we shall consider only these two contributions; at
next-to-leading order (which is the accuracy we need here) there are
also contributions of other operators in the effective Hamiltonian.
However, these contributions are small and can be neglected
here \cite{WiseLigeti}, although they should be included when
analyzing experimental data.
Considering the inclusive $B$ decays mediated by these two effective
interactions, we shall look at the generic quantity
\begin{equation}
R = \sum_X (2\pi)^4 \delta^4 (p_B - p_X - q)
\langle B(p_B)| \bar{b} \Gamma q | X (p_X) \rangle
\langle X (p_X)| \bar{q} \Gamma^\dagger b | B(p_B) \rangle
\end{equation}
where $q$ is the momentum transferred to the non-hadronic system and
$\Gamma$ is either $\gamma_\mu (1-\gamma_5)$ or $\sigma_{\mu \nu}$.
We first redefine the phase of the heavy quark field
\begin{equation}
b(x) = \exp(-im_b v \cdot x) b_v (x)
\end{equation}
where $v$ is the velocity of the decaying $B$ meson
\begin{equation} \label{velo}
v = \frac{p_B}{M_B} \, ;
\end{equation}
this corresponds to a splitting of the $b$ quark momentum according to
$p_b = m_b v + k$, where $k$ is a small residual momentum. Going through
the usual steps, we can rewrite $R$ as
\begin{equation}
R = \int d^4 x \exp(-ix[m_b v - q])
\langle B(p_B)| \bar{b}_v (0) \Gamma q(0)
\bar{q} (x) \Gamma^\dagger b_v(x) | B(p_B) \rangle
\end{equation}
The momentum $P=m_b v -q$ is the momentum of the final state partons,
which is considered to be large compared to to any of the
scales appearing in the matrix element. This allows us to set up
an operator product expansion (OPE).
If we assume
\begin{equation}
(m_b v - q)^2 = {\cal O}(m^2) \mbox{ and } (m_b - v q) = {\cal O}(m)
\end{equation}
the OPE is a short distance expansion yielding the usual $1/m_b$
expansion of the rates.
Close to the endpoint, where $P=m_b v - q$ is a practically
light-like vector, which has still large components, i.e.
\begin{equation} \label{kinema}
(m_b v - q)^2 = {\cal O}(\Lambda_{QCD} m_b)
\mbox{ and } (m_b - v q) = {\cal O}(m_b)
\end{equation}
one has to switch to a light cone expansion very similar to what
is known in deep inelastic scattering.
In order to study the latter case in some more detail, it is useful
to define light-cone vectors as
\begin{equation}
n_\pm = \frac{1}{\sqrt{(vP)^2 - P^2}} \left[
v \left(\sqrt{(vP)^2 - P^2} \mp (vP)\right) \pm P \right]
\quad n_\pm^2 = 0
\end{equation}
which satisfy the relations $n_+ n_- = 2$, $vn_\pm = 1$ and
$2v = n_+ + n_-$. These vectors allow us to
write
\begin{equation}
P = \frac{1}{2} ( P_+ n_- + P_- n_+ ) \qquad P_\pm = P n_\pm
\end{equation}
where
\begin{eqnarray}
&& P_+ = (vP) - \sqrt{(vP)^2 - P^2}
\to - \frac{P^2}{2 (vP)} \mbox { for } P^2 \ll (vP)^2 \\
&& P_- = (vP) + \sqrt{(vP)^2 - P^2}
\to 2 (vP) \mbox { for } P^2 \ll (vP)^2
\end{eqnarray}
The kinematic region in which the light cone distribution function
becomes relevant is the one where $P_+$ is much smaller than $P_-$:
Here the main contributions
to the integral come from the light cone $x^2 \approx 0$, since
in order to have a contribution to the integral we need to have
\begin{equation}
x\cdot P = \frac{1}{2} (x_- P_+ + x_+ P_-) = x_+ (vP) + \mbox{ const}
< \infty
\end{equation}
in the limit in which $ P_- \to \infty$. Hence $x_+$ has to be small,
restricting the integration to the light cone.
We consider first the tree level contribution,
which corresponds to a contraction of the light quark line;
we get \cite{MannelNeubert}
\begin{eqnarray}
R &=& \int d^4 x \int \frac{d^4 Q}{(2\pi)^4} \Theta(Q_0) (2\pi)
\delta(Q^2) \exp(-ix[m_b v - q - Q]) \\ \nonumber
&& \langle B(v)| \bar{b}_v (0) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q}
\Gamma^\dagger b_v(x) | B(v) \rangle
\end{eqnarray}
where now $|B(v)\rangle$ is the static $B$ meson state.
Performing a (gauge covariant) Taylor expansion of the
remaining $x$ dependence of the matrix element, we get
\begin{eqnarray} \label{taylor}
R &=& \int d^4 x \int \frac{d^4 Q}{(2\pi)^4} \Theta(Q_0) (2\pi)
\delta(Q^2) \exp(-ix[m_b v - q - Q]) \\ \nonumber
&& \sum_n \frac{1}{n!} \langle B(v)| \bar{b}_v (0) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q}
\Gamma^\dagger (-ix \cdot iD)^nb_v(0) | B(v) \rangle
\\ \nonumber
&=& \int d^4 x \int \frac{d^4 Q}{(2\pi)^4} \Theta(Q_0) (2\pi)
\delta(Q^2) \\ \nonumber
&& \langle B(v)| \bar{b}_v (0) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q}
\Gamma^\dagger {\cal P} \exp(-ix[m_b v - q - Q + iD])
b_v(0) | B(v) \rangle
\end{eqnarray}
where the symbol ${\cal P}$ means the usual path--ordering of the exponential.
Using spin symmetry and the usual representation matrices of the
$0^-$ B meson states,
the matrix element which appears in (\ref{taylor}) becomes
\begin{eqnarray}
&& \langle B(v)| \bar{b}_v (0) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q}
\Gamma^\dagger
(iD_{\mu_1}) (iD_{\mu_2}) \cdots (iD_{\mu_n}) b_v(0) | B(v) \rangle
\\ \nonumber
&& = \frac{M_B}{2} \mbox{Tr}\{\gamma_5 (\@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{v}+1) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q}
\Gamma^\dagger (\@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{v}+1)\gamma_5 \}
[a_1^{(n)} v_{\mu_1} v_{\mu_2} \cdots v_{\mu_n}
+ a_2^{(n)} g_{\mu_1 \mu_2} v_{\mu_3} \cdots v_{\mu_n} + \cdots ]
\end{eqnarray}
where the elipses denote terms with one or more $g_{\mu \nu}$'s and
also antisymmetric terms and $a_i^{(n)}$ are non-pertrubative parameters.
Contracting with
$x^{\mu_1} x^{\mu_2} \cdots x^{\mu_n}$ all antisymmetric contributions
vanish; furthermore, since the relevant kinematics restricts the $x_\mu$
to be on the light cone, also all the $g_{\mu \nu}$ terms are suppressed
relative to the first term, which has only $v_\mu$'s. Hence
\begin{equation} \label{term}
\langle B(v)| \bar{b}_v (0) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q}
\Gamma^\dagger (-ix \cdot iD)^n b_v(0) | B(v) \rangle
= M_B \mbox{Tr}\{(\@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{v}+1) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q} \Gamma^\dagger \}
a_1^{(n)} (v\cdot x)^n
\end{equation}
In this way we have made explicit that the kinematics we are studying
here forces us to resum the series in $1/m_b$. Defining the twist
$t$ of an operator ${\cal O}$ in the usual way
$t = dim[{\cal O}] - \ell$, where $\ell$ is the spin of the operator, we
find that the resummation corresponds to the contributions of leading
twist, $t=3$.
Since $x_\mu$ is light like, it projects out only the light cone
component $D_+$ of the covariant derivative in (\ref{term}).
Hence we may write $a_1^{(n)}$ as
\begin{equation}
2 M_B a_1^{(n)} = \langle B(v)| \bar{b}_v (iD_+)^n b_v | B(v) \rangle
\end{equation}
and one obtains as a final result
\begin{eqnarray}
R &=& \int d^4 x \int \frac{d^4 Q}{(2\pi)^4} \Theta(Q_0) (2\pi)
\delta(Q^2) \frac{1}{2}
\mbox{Tr}\{(\@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{v}+1) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q} \Gamma^\dagger \}
\\ \nonumber
&& \langle B(v)| \bar{b}_v \exp(-ix[(m_b+iD_+) v - q - Q])
b_v | B(v) \rangle
\end{eqnarray}
Introducing the shape function (or light cone distribution function)
as \cite{Neubert,MannelNeubert,Bigietal}
\begin{equation}
2 M_B f(k_+) = \langle B(v)| \bar{b}_v \delta(k_+ - iD_+) b_v | B(v) \rangle
\end{equation}
we can write the result as
\begin{eqnarray} \label{final}
R &=& \int dk_+ f(k_+) \int \frac{d^4 Q}{(2\pi)^4} \Theta(Q_0) (2\pi)
\delta(Q^2) \\ \nonumber && M_B
\mbox{Tr}\{(\@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{v}+1) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q} \Gamma^\dagger \}
(2 \pi)^4 \delta^4 ([m_b + k_+]v -Q - q)
\end{eqnarray}
This result shows that the leading twist contribution is obtained
by convoluting the partonic result with the shape function, where
in the partonic result the $b$ quark mass $m_b$ is replaced by
the mass
\begin{equation} \label{mstar}
m_b^* = m_b + k_+
\end{equation}
Let us take a closer look at the kinematics. The final-state quark
is massless,
$$
0 = ([m_b+k_+] v - q)^2
$$
in which case we have for
the light cone component of the heavy quark residual momentum\footnote{%
There are actually two solutions, but only one vanishes
at the endpoint}
\begin{equation} \label{kplus}
k_+ = -m_b + (vq) + \sqrt{(vq)^2 - q^2}
\end{equation}
Note that this variable -- up to a minus sign -- is just $P_+$ and
thus close to the endpoint
\begin{equation}
k_+ \approx - \frac{P^2}{2(vP)}
\end{equation}
If we now look at the hadronic kinematics
\begin{equation}
M_B v - q = p_X \, ,
\end{equation}
where $p_X$ is the four momentum of the final state hadrons,
and use the relation
between the heavy quark mass and the meson mass
\begin{equation} \label{mass}
M_B = m_b + \bar{\Lambda} + {\cal O}(1/m_b)
\end{equation}
we find
\begin{equation}
p_X = P + \bar{\Lambda} v \quad \mbox{ or } \quad
p_{X\pm} = P_\pm + \bar{\Lambda}
\end{equation}
Thus the light cone component
of the hadronic momentum of the final state
\begin{equation}
L_+ = - p_{X+} = - M_B + (vq) + \sqrt{(vq)^2 - q^2}
\end{equation}
ranging between $-M_B \le L_+ \le 0$ is directly related to
the light-cone component of the residual momentum of the heavy
quark
\begin{equation}
L_+ = k_+ - \bar{\Lambda}
\end{equation}
where the range of $k_+$ is given by $-m_b \le k_+ \le \bar{\Lambda}$.
However, large negative values of $k_+$ close to $-m_b$ are beyond
the validity of the heavy mass limit.
The observable $L_+$ directly measures the light cone component
of the residual momentum of the heavy quark, at least for small values
of $L_+$. In the case of $b \to s \gamma$ we have
$q^2 = 0$ and $L_+$ is directly related to the energy $E_\gamma$ of the photon
in the rest frame of the decaying $B$
$$
L_+ = - M_B + 2 E_\gamma .
$$
For
$b \to u \ell \bar{\nu}_\ell$ a reconstruction of $L_+$ requires both
a measurement of the hadronic energy and the hadronic invariant mass.
Reexpressing the
convolution (\ref{final}) in terms of $L_+$ and using that the
light-cone distribution function is non-vanishing only for
$-m_b \to - \infty < k_+ < \bar{\Lambda}$
we may rewrite (\ref{final}) as
\begin{eqnarray} \label{final1}
R &=& \int\limits_{-M_B}^0 dL_+ f(L_+ +\bar{\Lambda})
\int \frac{d^4 Q}{(2\pi)^4} \Theta(Q_0) (2\pi)
\delta(Q^2) \\ \nonumber && M_B
\mbox{Tr}\{(\@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{v}+1) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q} \Gamma^\dagger \}
(2 \pi)^4 \delta^4 ([M_B + L_+]v -Q - q)
\end{eqnarray}
where we have made use of (\ref{mass}). Equivalently, we may write
the spectrum in the variable $L_+$ as
\begin{eqnarray} \label{final2}
\frac{dR}{dL_+} &=& f(L_++\bar{\Lambda})
\int \frac{d^4 Q}{(2\pi)^4} \Theta(Q_0) (2\pi)
\delta(Q^2) \\ \nonumber && M_B
\mbox{Tr}\{(\@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{v}+1) \Gamma \@ifnextchar[{\fmsl@sh}{\fmsl@sh[0mu]}{Q} \Gamma^\dagger \}
(2 \pi)^4 \delta^4 ([M_B + L_+]v -Q - q)
\end{eqnarray}
The two relations (\ref{final1}) and (\ref{final2}) exhibit an
interesting feature of the summation of the leading twist terms, namely
that the dependence on the heavy quark mass has completely disappeared,
only the shape function $f$ still depends on $\bar\Lambda$. The full result
should be independent of this unphysical quantity and hence the
explicit dependence on $\bar{\Lambda}$ of the shape
function has to cancel against the one appearing in the argument of the
shape function. In other words, if we write the shape function as
$f = f(k_+,\bar{\Lambda})$, this function of two variables has to satisfy
\begin{equation} \label{Labindependence}
\left(\frac{\partial}{\partial k_+} +
\frac{\partial}{\partial \bar{\Lambda}} \right) f(k_+,\bar{\Lambda})
= 0 \quad \mbox{ or } \quad f(k_+,\bar{\Lambda}) = g(k_+ - \bar{\Lambda})
\end{equation}
In this way the total result becomes independent of any reference to the
quark mass $m_b$ or equivalently on $\bar{\Lambda}$. In particular,
(\ref{Labindependence}) ensures the cancellation
of ambiguities related to the renormalon in the heavy quark mass
\cite{BigiUraltsev}.
Putting everything together, the
light-cone spectrum (\ref{final2}) for the decay
$B \to X_u \ell \bar{\nu}_\ell$ becomes
\begin{equation} \label{final3}
\frac{d \Gamma}{dL_+} (B \to X_u \ell \bar{\nu}_\ell) =
\frac{G_F^2 (M_B + L_+)^5 |V_{ub}|^2}{192 \pi^3} g(L_+)
\end{equation}
which means that the spectrum is proportional to the total rate
in which the $m_b$ has been replaced by $M_B + K_+$.
Similarly, for $B \to X_s \gamma$ one finds
\begin{equation} \label{final4}
\frac{d \Gamma}{dL_+} (B \to X_s \gamma)=
\frac{G_F^2 (M_B + L_+)^5 |V_{tb} V_{ts}^*|^2 \alpha_{em} }{32 \pi^4}
[C_7 (M_B)]^2 g(L_+)
\end{equation}
Note that (\ref{final3}) and (\ref{final4}) do not depend on
any unphysical parameter,
since the $\bar\Lambda$ dependence has disappeared. In particular,
all ambiguities induced by renormalons should cancel in these equations.
Based on this it has been suggested to give a definition for a
heavy quark mass $\hat{m}_b$ through (\ref{final3}) and (\ref{final4})
using e.g. the mean energy $\langle E_\gamma \rangle$ of the photon in
$B \to X_s \gamma$ \cite{Bauer}. To this end we consider
the moments of the function $g$, which can be obtained from the
moments of the shape function $f$. The usual HQET relations are
\begin{eqnarray} \label{moments}
&& \int_{-\infty}^{\bar\Lambda} dk_+ f(k_+) = 1 \label{zero}\\
&& \int_{-\infty}^{\bar\Lambda} dk_+ k_+ f(k_+) = \delta m_b \label{one}\\
&& \int_{-\infty}^{\bar\Lambda} dk_+ k_+^2 f(k_+) \label{two}
= -\frac{1}{3} \lambda_1
\end{eqnarray}
which define the normalization, the residual mass term \cite{FalkNeubert}
and the kinetic energy parameter.
Reexpressing relation (\ref{one}) through the function $g$ we have
\begin{equation}
\int_{-\infty}^0 dL_+ L_+ g(L_+) = \delta m_b - \bar\Lambda
\equiv \hat\Lambda
\end{equation}
which gives a definition of the heavy quark mass
$\hat{m_b} = M_B - \hat\Lambda$ free of renomalon ambiguities which
cancel between the $\delta m_b$ and $\bar\Lambda$ \cite{Bauer}.
For the second moment one obtains from (\ref{two})
\begin{equation} \label{two1}
\int_{-\infty}^0 dL_+ L_+^2 g(L_+)
= -\frac{1}{3} \lambda_1 - \bar\Lambda^2 - 2 \bar\Lambda \hat\Lambda
\equiv -\frac{1}{3} \hat\lambda_1
\end{equation}
where $\hat\lambda_1$ again has to be free of renormalon ambiguities.
We shall later need a model shape function to discuss our results.
In order to keep things simple, we use the one-parameter
model of \cite{MannelNeubert}
\begin{eqnarray} \label{sf}
g(L_+) &=& {32 \over \pi^2 \sigma^3} L_+^2
\exp\left[- {4 \over \pi \sigma^2 } L_+^2 \right]
\Theta(-L_+)
\end{eqnarray}
which is easy to deal with. Here $\sigma$ is the width
parameter of the shape function and will be varied in our numerical
studies between 400 MeV and 800 MeV.
\section{Including Perturbative Contributions}
The next step is to include radiative corrections to the relations
(\ref{final3}) and (\ref{final4}). The starting point of the
considerations is (\ref{final}) or eqivalently (\ref{final1}).
However, now the partonic result $d\Gamma_{part}$ including
the radiative corrections is convoluted with the shape
function
\begin{equation}
d\Gamma_{had} = \int\limits_{-m_b}^{\bar{\Lambda}}
d\Gamma_{part} (m_b^* = m_b + k_+) f(k_+) dk_+
\end{equation}
We shall compute the spectrum in the light cone variable of the
hadronic momentum $L_+$, and the partonic counterpart of this
variable is
\begin{equation}
l_+ = - m_b + (vq) + \sqrt{(vq)^2 - q^2} \, .
\end{equation}
in which we compute a differential spectrum as some function $G$ of
$l_+$ and $m_b$
\begin{equation}
\frac{d\Gamma_{part}}{d l_+} = G(m_b,l_+) \, .
\end{equation}
To apply the convolution formulae (\ref{final}) and (\ref{final1})
we replace in the first step the heavy quark mass by $m_b^*$
and convolute with the shape function. The variable $l_+$
is also replaced by
\begin{equation}
l_+^* = - m_b^* + (vq) + \sqrt{(vq)^2 - q^2} = l_+ - k_+
\end{equation}
Note that $l_+^*$ is again a light cone variable for a process in which
the heavy quark mass is replaced by $m_b^*$, and hence it ranges between
$-m_b^* \le l_+^* \le 0$, which means that the $k_+$ integration becomes
restricted to
\begin{equation}
-l_+ \le k_+ \le \bar{\Lambda}
\end{equation}
The second step towards hadronic variables is to replace $l_+$ by $L_+$,
the hadronic light cone momentum. At the same time we perform a shift
in the integration variable $k_+ \to K_+ - \bar{\Lambda}$ and obtain
\begin{equation} \label{pert}
\frac{d\Gamma_{had}}{dL_+} = \int\limits_{L_+}^0 dK_+ \,
G(M_B+K_+,L_+ -K_+) g(K_+)
\end{equation}
Relation (\ref{pert}) holds for any $b \to q$ transition where $q$ is
a light quark. We shall exploit (\ref{pert}) for a comparison between
the inclusive processes $B \to X_s \gamma$ and
$B \to X_u \ell \bar{\nu}_\ell$. The partonic $l_+$ spectra of the two
processes can be calculated; both take the generic form
\begin{eqnarray} \label{partonic}
G(m_b,l_+) &=& \gamma_0 m_b^5
\left[ (1+ b_0 \frac{\alpha_s}{3\pi}) \delta(l_+) +
\vphantom{\int\limits_t^t} \right. \\ \nonumber
&& \left. \frac{\alpha_s}{3\pi} \left(
b_1 \left(\frac{\ln(-l_+ / m_b)}{-l_+} \right)_+ +
b_2 \left(\frac{1}{-l_+} \right)_+
+ \frac{1}{m_b} \Phi \left(\frac{-l_+}{m_b} \right) \right) \right]
\end{eqnarray}
where we have defined the ``+''-distributions in the usual way
\begin{equation}
\int_0^1 dx D_+ (x) = 0 \, ,
\end{equation}
and $b_0$, $b_1$, $b_2$ and $\Phi$ are known quantities.
For $b\to u \ell \bar{\nu}_\ell$ we find
\begin{eqnarray} \label{b2uln}
\gamma_0 &=& {G_F^2 \left|V_{ub}\right|^2 \over 192 \pi^3} \\
b_0 &=& -{13\over 36}-2\pi^2 \nonumber \\
b_1 &=& -4 \nonumber\\
b_2 &=& -26/3 \nonumber \\
\Phi(x) &=& \frac{158}{9} + \frac{407\,x}{18} - \frac{367\,{x^2}}{6} +
\frac{118\,{x^3}}{3}
- \frac{100\,{x^4}}{9} + \frac{11\,{x^5}}{6} -
\frac{7\,{x^6}}{18} \nonumber \\
&& - \frac{4\,\ln (x)}{3} + \frac{46\,x\,\ln (x)}{3} +
6\,{x^2}\,\ln (x)
- \frac{16\,{x^3}\,\ln (x)}{3} \nonumber\\
&& - 12\,{x^2}\,{{\ln (x)}^2} + 8\,{x^3}\,{{\ln (x)}^2}
\nonumber
\end{eqnarray}
While for $b\to s\gamma$
\begin{eqnarray} \label{b2sg}
\gamma_0 &=& {G_F^2 \left|V_{ts}V_{tb}^*\right|^2 \left| C_7\right|^2
\alpha_{\rm em} \over 32 \pi^4} \\
b_0 &=& -5-\frac{4}{3} \pi^2 \nonumber \\
b_1 &=& -4 \nonumber \\
b_2 &=& -7 \nonumber \\
\Phi(x) &=& 6 + 3\,x - 2\,{x^2} - 4\,\ln (x) + 2\,x\,\ln (x) \nonumber
\end{eqnarray}
where $C_7 = -0.306$ is obtained from the next-to-leading order calculation
of \cite{ChetMM}. Our result for $B \to X_s \gamma$ coincides with the ones
obtained in the literature \cite{AliGreub,ChetMM,KaganNeubert}.
To determine $b_0$ for $B \to X_u \ell \bar{\nu}_\ell$ and $B \to X_s \gamma$
the respective total rates at ${\cal O}(\alpha_s)$ \cite{Timo,Pott} have
been used.
The partonic result up to order $\alpha_s$ contains terms of all
orders in the $1/m_b$ expansion and we shall first consider the
leading twist contribution. Taking the
limit $m_b \to \infty$ of the partonic calculation leaves us only
with the ``+''-distributions and the $\delta$ function. The
convolution becomes
\begin{eqnarray}
\frac{d\Gamma_{had}}{dL_+} &=& \int\limits_{L_+}^0 dK_+ \, g(K_+)
\gamma_0 (M_B+K_+)^5
\left[ (1+ b_0 \frac{\alpha_s}{3\pi}) \delta(L_+-K_+)
\vphantom{\int\limits_t^t} \right. \\ \nonumber
&& + \frac{\alpha_s}{3\pi} \left. \left(
b_1 \left(
\frac{\ln\left(\frac{-L_+ + K_+}{M_B + K_+}\right)}{-L_+ + K_+} \right)_+ +
b_2 \left(\frac{1}{-L_+ + K_+} \right)_+ \right) \right]
\end{eqnarray}
The shape function $g(K_+)$ is restricted to values $K_+$ small
compared to $M_B$, and one may expand the dependence on $M_B + K_+$
for small $K_+$. This induces terms of higher orders in $1/m_b$, which
again may be dropped. Hence the leading twist terms become
\begin{eqnarray}
\frac{d\Gamma_{had}}{dL_+} &=& \gamma_0 M_B^5
\left[ (1+ b_0 \frac{\alpha_s}{3\pi}) g(L_+) +
\vphantom{\int\limits_t^t} \right. \\ \nonumber
&& \left. \frac{\alpha_s}{3\pi} \int\limits_{L_+}^0 dK_+ g(K_+) \left(
b_1 \left(
\frac{\ln\left(\frac{-L_+ + K_+}{M_B}\right)}{-L_+ + K_+} \right)_+ +
b_2 \left(\frac{1}{-L_+ + K_+} \right)_+ \right) \right]
\end{eqnarray}
It is well known that the coefficient $b_1$ is universal as well as
the shape function.
This suggests to define a scale dependent shape function, which to
order $\alpha_s$ becomes
\begin{equation} \label{scale}
{\cal F}(K_+,\mu) = g(K_+) + \frac{\alpha_s b_1}{3 \pi}
\int\limits_{L_+}^0 dK_+ g(K_+) \left(
\frac{\ln\left(\frac{-L_+ + K_+}{\mu}\right)}{-L_+ + K_+} \right)_+
\end{equation}
The scale dependence of the distribution function has been discussed
in \cite{BMK}, where the evolution equation for the distribution function
has been set up.
The terms proportional to $(1/l_+)_+$ are not universal,
which is natural, since the relavant scale in two different processes
may be different. A change of scale changes the contribution
of the $(\alpha_s / \pi) (1/l_+)_+$ pieces according to
\begin{equation}
{\cal F}(K_+,\mu) = {\cal F}(K_+,\mu^\prime) +
\frac{\alpha_s b_1}{3 \pi}
\int\limits_{L_+}^0 dK_+ g(K_+)
\ln\left(\frac{\mu^\prime}{\mu}\right)
\left(\frac{1}{-L_+ + K_+} \right)_+
\end{equation}
and hence we find for the leading twist contribution
\begin{equation} \label{part1}
\frac{d\Gamma_{had}}{dL_+} = \gamma_0 M_B^5
(1+ b_0 \frac{\alpha_s}{3\pi}) {\cal F}(L_+,\mu)
\end{equation}
where the scale $\mu$ is given by
\begin{equation} \label{scalemu}
\ln\left(\frac{M_B}{\mu}\right) = \frac{b_2}{b_1}
\end{equation}
Inserting the results for the two processes under consideration
we find
\begin{equation} \label{scales}
\mu_{b \to s \gamma} = 0.1738 M_b \qquad
\mu_{b \to u \ell \bar{\nu}_\ell} = 0.1146 M_b
\end{equation}
which are in fact scales small compared to the mass of the $b$ quark.
Furthermore, the ratio of the two scale is of order one,
\begin{equation}
\frac{\mu_{b \to s \gamma}}{\mu_{b \to u \ell \bar{\nu}_\ell}} \approx
1.52
\end{equation}
The physical meaning of the scale $\mu$ can be understood in a picture
as proposed in \cite{KorchemskySterman}. Here the amplitude is decomposed
into a hard part, a ``jetlike'' part incorporating the collinear
singularities and a soft part. The typical scales
corresponding to these pieces
are $m_b^2$ for the hard, $\Lambda_{QCD} m_b$ for the ``jetlike'' and
$\Lambda_{QCD}^2$ for the soft part. Comparing the present approach
to \cite{KorchemskySterman} the light-cone distribution function
corresponds to the convolution of the soft and the ``jetlike'' piece,
both of which are universal but scale dependent.
Thus the scale appearing in the shape function should be of the order
$\mu^2 \sim m_b \Lambda_{QCD}$ since it incorporates all lower scales.
In particular, the shape function combines both the collinear and the
soft contributions which according to \cite{KorchemskySterman} could be
factorized. Since we are working to
next--to--leading order, we can fix the scale according to (\ref{scalemu}),
and we expect to obtain a scale $\mu^2$ of order
$\Lambda_{QCD} m_b \sim 0.9$ GeV${}^2$, which is confirmed by (\ref{scales}).
The rest of the partonic result, i.e. the function $\Phi$,
is a contribution of subleading terms, suppressed by at least one
power of the heavy mass. We shall use these terms to estimate, how
far we can trust the leading twist terms, in particular in the
comparison between
$B \to X_u \ell \bar{\nu}_\ell$ and $B \to X_s \gamma$.
To this end we add the two contributions and use the expression
\begin{equation} \label{partfull}
\frac{d\Gamma_{had}}{dL_+} = \gamma_0 M_B^5 \left[
(1+ b_0 \frac{\alpha_s}{3\pi}) {\cal F}(L_+,\mu)
+ \frac{\alpha_s}{3\pi M_B} \Phi \left(-\frac{L_+}{M_B} \right) \right]
\end{equation}
We note that the functions $\Phi$ for both processes contain terms
diverging logarithmically as $L_+ \to 0$, but which are suppressed
by $1/m_b$. These divergencies will be cured once the analogon of the
light-cone distribution function at subleading order is taken into
account. However, for the quantitative analyis of the process these
terms are not important as long as we do not get too close to the
endpoint.
\section{Comparison of $B \to X_u \ell \bar{\nu}_\ell$
to $B \to X_s \gamma$}
\begin{figure}
\hspace*{-15mm}
\leavevmode
\epsfxsize=8cm
\epsffile{plot1a.eps}
\epsfxsize=8cm
\epsffile{plot1b.eps}
\caption{Light-cone distribution spectra for
$B \to X_u \ell \bar{\nu}_\ell$ (solid) and
$B \to X_s \gamma $ (dashed) in units of $\gamma_0 M_B^5$.
The dotted line is the leading contribution, which is the
same for both processes. The three sets of curves correspond
to values of $\sigma$ of 400, 600 and 800 MeV.}
\label{cfig1}
\end{figure}
We shall first study the $L_+$ spectra of the two processes
using (\ref{part1}) and (\ref{partfull}). These results depend on the
model for the shape function one is using and we shall employ the simple
one-parameter model (\ref{sf}).
In fig.\ref{cfig1} we show the $L_+$ spectra of
$B \to X_u \ell \bar{\nu}_\ell$ and $B \to X_s \gamma$. We divide the
rates by the prefactors $\gamma_0 M_B^2$, such that at leading order
only the shape function remains. The left plot shows the shape
function (\ref{sf}) together with the
leading twist contribution (\ref{part1}) for the two processes
for three different values of the width parameter $\sigma$.
The leading twist contribution can only be trusted close to the endpoint
$L_+ =0$; it even becomes negative for values
$L_+ \le - \sigma$, indicating the
breakdown of the leading twist approximation. Adding the subleading
terms as in (\ref{partfull}) cures this problem and the results are
shown in the right hand figure of fig.\ref{cfig1}. The sharp rise of the
full results is due to the logarithms $\ln L_+$ of order $1/m_b$, which
become relevant only very close to $L_+ = 0$. These contributions are
integrable and will not be relevant once the rates are binned with
reasonably large bins.
Next we shall compare $B \to X_u \ell \bar{\nu}_\ell$
to $B \to X_s \gamma$ by considering the ratio of the two $L_+$
distributions. From the experimental point of view many systematic
uncertainties cancel, while from the theoretical side one
expects to reduce nonperturbative uncertainties, which indeed
at tree level cancel completely.
The perturbative result for the spectrum is a distribution and
the ratio of the perturbative expressions is meaningless, even if
we avoid the region around $L_+ = 0$. However, once we have combined
both perturbative and non-perturbative contributions we can take the
ratio without encountering the problem of distributions.
The price we have to pay is that the resulting ratio is not
independent of the shape function, an effect which has
been observed already in \cite{Neubert1}. Still the ratio of the
two processes is a useful quantity, since one may expect that
much of the non-perturbative uncertainties still cancel.
\begin{figure}
\hspace*{-15mm}
\leavevmode
\epsfxsize=8cm
\epsffile{plot2a.eps}
\epsfxsize=8cm
\epsffile{plot2b.eps}
\caption{Ratio of the light cone distribution spectra for the two
processes. The leading twist contribution (\protect{\ref{comp}})
is given in the left plot while the right plot shows
the full result.
The solid, dashed and dotted lines correspond to values
of $\sigma$ of 600, 400 and 800 MeV, respectively.}
\label{cfig2}
\end{figure}
The two spectra are due to the ``smearing'' with the nonperturbative
shape function smooth curves and we can perform a comparison of the
two processes by taking the ratio. At leading twist this ratio becomes
\begin{equation} \label{comp}
\frac{(d \Gamma_{B \to X_u \ell \bar{\nu}_\ell} / dL_+)}
{(d \Gamma_{B \to X_s \gamma} / dL_+)}
= \frac{|V_{ub}|^2}{|V_{ts} V_{tb}^*|^2}\frac{\pi}{6 \alpha_{em} |C_7|^2}
\frac{{\cal F}(L_+,\mu_{b \to s \ell \bar{\nu}_\ell})}
{{\cal F}(L_+,\mu_{b \to s \gamma})}
\left(1 + \frac{\alpha_s (m_b)}{3 \pi} \left[\frac{167}{36} -
\frac{2}{3} \pi^2 \right] \right)
\end{equation}
where the scales of the distribution functions have been given
in (\ref{scales}). It is interesting to note that although the radiative
corrections at leading twist (the constants $b_0$) are big for the
individual processes, they turn out to be small in the ratio (\ref{comp}).
For values of $L_+$ within the nonperturbative region
$L_+ \ge -\sigma$
the leading twist contribution is dominant and the comparison may be
performed using the leading twist terms only. This is the region with
the largest fraction of the total rate and hence the subleading terms
will not play a role, also due to experimental cuts. The ratio of the
leading twist terms is shown in the left hand plot of fig.\ref{cfig2}
for three values of the parameter $\bar\Lambda$.
The region of validity of the leading twist contribution may be studied
by looking at the ratio of the full results (\ref{partfull}) shown in
fig(\ref{cfig2}). The subleading terms lead to a minimum at the point
where they are becomming the dominant contribution to the rates. This
happens at values of $L_+$ which correspond to the typical width $\sigma$
of the non-perturbative shape function.
For values of $L_+$ between $-\sigma$ and $0$ one may then use our
results to determine CKM matrix elements, since the ratio of the
two rates depends on
$$
\frac{|V_{ub}|^2}{|V_{ts} V_{tb}^*|^2}\frac{\pi}{6 \alpha_{em} |C_7|^2}
\approx 5
$$
i.e. the results shown in fig.\ref{cfig2} have to be multiplied by
this factor. Hence one may determine $ |V_{ub}|/|V_{ts}|$ in this
way; however, the radiative corrections induce a small hadronic
uncertainty, which is hard to quantify without knowledge of the
non-perturbative distribution function. Still one may expect only a
small uncertainty, since at tree level there is no hadronic uncertainty
at all.
\section{Conclusions}
Due to experimental constraints severe cuts on the phase space in
both $b \to s \gamma$ and $b \to u \ell \bar{\nu}_\ell$ inclusive
transitions have to be imposed making these decays sensitive to
non-perturbative effects. Within the framework of the heavy mass
expansion of QCD these effects are encoded in a light-cone distribution
function which is universal for all heavy to light processes.
The universality of this function may be exploited to eliminate
the non-perturbative uncertainties in $b \to s \gamma$ using the
input of $b \to u \ell \bar{\nu}_\ell$ or vice versa. To this end,
one may use the comparison of the two processes to determine the
ratio $V_{ub} / V_{ts}$ or --- for fixed CKM parameters --- to test
the Wilson coefficient $C_7$.
The comparison between the two processes is best performed in terms
of the spectra of the light-cone component of the final state hadrons,
which for the case of $B \to X_s \gamma$ is equivalent to
the photon energy spectrum, while for $B \to X_u \ell \bar{\nu}_\ell$
this requires a measurement of both the hadronic invariant mass and the
hadronic energy.
For small values, the light cone component of the final state hadrons is
the same as the light cone component of the heavy quark residual momentum.
The corresponding spectra are at tree level directly proportional to
the light cone distribution function and hence one may access this
function by a measurement of the the light cone distribution
of the momentum of the final state hadrons.
However, radiative corrections change this picture. The main effect is
that the shape function does not cancel any more in the ratio of the
two spectra. The
partonic radiative corrections to the light-cone distribution of
the final state partons exhibit the well known singularities of
the type $\ln l_+ / l_+$ and $1/l_+$ which can be absorbed into
a radiatively corrected, scale dependent light cone distribution function.
It turns out that the relevant scales are small in both cases and
different. The scale dependent distribution functions involve a
convolution such that the nonperturbative effects cannot be cancelled
anymore in the ratio.
We have also computed the subleading terms partonically and use
this to estimate the reliability of the leading twist comparison of
$B \to X_s \gamma$ and $B \to X_u \ell \bar{\nu}_\ell$. This
comparison depends on the width of the non-perturbative
function once radiative corrections are taken into account.
For values of $L_+$ close to the endpoint (i.e. within the width
of the distribution function) one may use the leading twist
contribution to extract $|V_{ub}| / |V_{ts}|$ with reduced hadronic
uncertainties. This method should become feasible at the future
$B$ physics experiments.
\section*{Acknowledgements}
We thank Kostja Chetyrkin, Timo van Ritbergen and Marek Jezabek for
discussions and comments. SR and TM acknowledge the support of the DFG
Graduiertenkolleg ``Elementarteilchenphysik an Beschleunigern'';
TM acknowledges the support of the DFG Forschergruppe
``Quantenfeldtheorie. Computeralgebra und MonteCarlo Simulationen''.
|
\section{Introduction}
The ordering of holes and spins into stripes in the CuO$_2$-planes of
Nd-doped La$_{2-x}$Sr$_{x}$CuO$_4$ (LSCO) has attracted much attention, because
charge separation and antiferromagnetic spin fluctuations are believed to
be important
for understanding the mechanism of high temperature superconductivity.
The idea of pinned stripes in the low temperature tetragonal (LTT) phase
of Nd-doped LSCO or La$_{2-x}$Ba$_{x}$CuO$_4$ \cite{john_nature,axe89} provides
an explanation for the anomaly observed at $x \sim \frac{1}{8}$, where
superconductivity is destroyed, or at least strongly suppressed
\cite{mood88,craw91,buch94,mood97}. The tilting of the oxygen octahedra
along the $[1\,0\,0]$ and $[0\,1\,0]$ directions (i.e., parallel to Cu--O
bonds) in
the LTT phase introduces a pinning potential for horizontal and vertical
stripes,
whereas in the low temperature orthorhombic (LTO) phase it is absent due to the
rotation of the tilting axis into the $[1\,1\,0]$ direction.
In a single crystal of La$_{1.48}$Nd$_{0.4}$Sr$_{0.12}$CuO$_4$, neutron
diffraction
allowed the observation of both magnetic and charge-ordering superstructure
reflections \cite{john_nature}. Consistent with the idea of pinning by the LTT
lattice modulation, static stripe ordering within the CuO$_2$ planes
appears at the
transition temperature from the LTO to the LTT phase, which is $\rm \sim 68 K$.
From the positions of the superlattice peaks and the nominal hole
concentration it
follows that stripes of holes are approximately half-filled, and act as
antiphase domain walls with respect to the antiferromagnetically ordered Cu
spins. Thus, a Sr doping level of $\sim\frac{1}{8}$ yields a spacing between
stripes of $4a$, where $a$ is the lattice constant.
High energy x-ray studies have been succesful in confirming the results for the
charge stripe ordering in an $x=0.12$ sample \cite{martin98}.
It is still an open question as to whether charge
stripes are limited to hole concentrations near $\frac{1}{8}$ in Nd-doped
LSCO or
whether they influence the physics of cuprate superconductivity in general,
but there is accumulating evidence for the latter.
Inelastic neutron scattering experiments indicate the possibility of
moving, fluctuating stripes in LSCO \cite{cheo91,yamada98} and even in YBCO
\cite{dai98,mook98}. Local charge ordering in
La$_{2-x}$Sr$_{x}$CuO$_4$ with $x\leq\frac18$ is suggested by a recent
nuclear quadrupole resonance (NQR) study \cite{hunt99}.
It is interesting to imagine that dynamic stripe correlations might be
necessary for
superconductivity \cite{emer97}, whereas pinning the stripes leads to a
strongly
reduced critical temperature $\rm T_c$, but the experimental evidence for
such a
scenario is still incomplete.
So far the Nd-doped LSCO system offers a unique
opportunity to study pinned stripe patterns in diffraction experiments,
giving information about the nature of charge and spin ordering.
The superstructure reflections due to the ordering of holes are shifted by
$2\epsilon$ in $h$- or $k$-direction relative to fundamental
reflections, whereas the magnetic peaks are located around the
antiferromagnetic peak position $(\frac12,\frac12,0)$, shifted by
$\epsilon$ also in $h$- or $k$-direction.
In neutron scattering experiments, the splitting $\epsilon$ has been
observed to
increase slightly as the Sr content increases from 0.12 to 0.15 and 0.20,
implying
a decrease in the average stripe spacing; however, it has only been
practical to
study the superstructure reflections due to the antiferromagnetic ordering
\cite{john97}. It is now imperative to directly characterize the charge
order at
Sr concentrations away from $\frac18$.
In this work we present x-ray studies with 120~keV photons on
$\rm La_{1.45}Nd_{0.4}Sr_{0.15}CuO_4$. Our results are in complete
agreement with the stripe model and complementary to the
experimental evidence of
antiferromagnetic stripe ordering in this sample \cite{john97}.
As observed in the $x=0.12$ case \cite{john_nature}, the superstructure
reflections due to the ordering of the holes sets in at a higher temperature
compared to the magnetic signal, indicating that the transition into
the stripe phase is driven by the charge separation \cite{zach98}.
\section{Experiment}
The experiments have been performed on the triple-axis diffractometer
designed for the use of $\sim 100$~keV photons
at the high-field wiggler beamline
BW5 at HASYLAB, Hamburg \cite{bouchard98}.
X-ray diffraction in this energy range has proven to be very
successful in studying charge ordering in cuprates, nickelates and manganates
\cite{martin98,titi97,thomas99}.
As in neutron scattering experiments, the large penetration depth
($\sim 1$ mm) allows one to probe the bulk of the sample, enabling a direct
comparison of x-ray and neutron diffraction data.
In contrast to former experiments on $\rm La_{1.48}Nd_{0.4}Sr_{0.12}CuO_4$
\cite{martin98}, the Si/TaSi$_2$ monochromator and analyzer crystals
have been replaced
by the new Si$_{1-x}$Ge$_x$ gradient crystal material. These crystals show
very high reflectivity values of 96\%\ (not corrected for absorption) and
variable widths of the rocking curves depending on the Ge content
\cite{steffen98,steffen99}.
By use of this new material,
the scattered intensity at the stripe peak positions
in Nd doped LSCO is about four times higher
compared to the results obtained previously with the
utilization of $\rm Si/TaSi_2$
crystals as monochromator and analyzer \cite{steffen98user}.
Figure~1 shows a comparison of a superstructure reflection measured with the
$\rm Si/TaSi_2$ and Si$_{1-x}$Ge$_x$ gradient crystals. The count rate
collected with the gradient crystals is nearly two times higher, and a
slightly better signal to background ratio could be reached by reducing the
beam spot on the sample, i.e. by probing only the center of the crystal.
The lower curve has
been measured with the $\rm Si/TaSi_2$ crystals, illuminating a two times
larger sample volume.
In this experiment 120~keV photons have been employed, with a monochromatic
beam intensity of $1.2\times 10^{11}$ photons/$\rm mm^2$.
The resolution (FWHM) at a (2,0,0) reflection of
$\rm La_{1.45}Nd_{0.4}Sr_{0.15}CuO_4$ was 0.020~\AA$^{-1}$ in the
longitudinal and 0.0014~\AA$^{-1}$ in the transverse direction, the latter
being
limited by the sample mosaicity. At smaller diffraction angles the longitudinal
resolution improves because the diffraction geometry becomes less
dispersive, e.g. the FWHM of the (1,1,0) reflection is
0.011~\AA$^{-1}$.
A closed-cycle cryostat has been utilised and temperatures between 9~K and
300~K could be reached at the sample position. The studied crystal
($\sim2\times2\times4$ mm$^3$) is a piece of a cylindrical rod that was
grown by the travelling-solvent floating-zone method.
\section{Results}
Previous hard x-ray diffraction experiments on a
$\rm La_{1.48}\-Nd_{0.4}Sr_{0.12}CuO_4$ crystal have
shown that the stripe peaks are displaced not only within the
$(h,k,0)$ plane, but also in the $\ell$ direction \cite{martin98}.
Due to the better resolution in reciprocal space compared to neutron
scattering experiments it has been possible to
find a modulated intensity of the stripe peaks
along the $\ell$-direction with maxima at
$\ell = \pm 0.5$, indicating a correlation of the charge stripes in
next-nearest-neighbor layers along the $c$-axis.
The highest scattering intensities for the stripe peaks
are expected at the positions $(2-2\epsilon,0,\pm0.5)$ and
$(2+2\epsilon,0,\pm0.5)$. Figure 2 clarifies the positions of the
stripe peaks in reciprocal space relative to the CuO$_2$ planes. We have used
the tetragonal unit cell with $a=b=3.775$~\AA, $c=13.10$~\AA.
Longitudinal scans along $(h,0,0.5)$ are shown in Fig.~3. As anticipated,
small superstructure reflections are observed at $h=2\pm2\epsilon$ with
$2\epsilon = 0.256(1)$. The peaks that are present at $T=9$~K have disappeared
after raising the temperature to 70~K. The signal to
background ratio is $\sim 0.2$ for the $2-2\epsilon$ reflection and $\sim 0.1$
at the $2+2\epsilon$ position.
In both scans the background rises due to the vicinity of the (2,0,0)
reflection. At this fundamental Bragg reflection
the count rate in the peak maximum is about $10^8$ times larger
than in the stripe peaks.
The curves through the data points in Fig.~3 are least-squares fits. In
both (a)
and (b) the charge-order peaks are modelled with a Gaussian. The
background in (a)
is approximated by the tail of a second Gaussian centered near $h=2$ plus a
linear
contribution, while only a linear background is used in (b).
Usually, the tail of a
fundamental reflection is Lorentzian shaped, but nevertheless it is
possible that locally other functions are a better approximation to a
non-linear background, since the shape of the
background is related to the sample
quality. The studied $\rm La_2CuO_4$ crystal with Sr and Nd as dopants
on the La-site certainly incorporates defects and strain which are
responsible for the rather high background, which is $\sim 150$ photons
per second at the superlattice positions.
In the inset of Fig.~3(a), two more low-temperature longitudinal scans,
slightly shifted in the $k$-direction, are displayed. The absence of any
peaks in
these scans indicates that the peak found at $(2-2\epsilon,0,0.5)$
is narrow in $k$. This conclusion is confirmed by the 9~K transverse
scan shown in Fig.~4. The background in the transverse scans is linear, and a
fit of the stripe peak with a Lorentzian is slightly more successful
than utilizing a Gaussian. From the present data, it is difficult to
determine the true peak shape, given the small signal-to-background ratio.
Figure~5(a) shows the temperature variation of the amplitudes of the $h$ and
$k$ scans at the $(2-2\epsilon,0,0.5)$ position.
With rising temperature the amplitudes decrease rather linearly; the
peak vanishes at 62(5)~K, which is $\sim8$~K below the structural
transition from
the LTT into the LTO phase. The intensity of the (3,0,0) reflection,
also shown in Fig.~5(a), is a measure of this
transition because it only occurs in the LTT phase.
From the longitudinal and transverse scans a FWHM of 0.028(5)~r.l.u.
can be inferred, resulting in a correlation length of 43(8)~{\AA}.
The change of the peak widths with temperature can be seen
in Fig.~5(b). Up to $\sim55$~K it is rather constant, but above this
temperature the width of the peak in $h$- and $k$-directions increases
considerably. Note that the increasing width as well as the
vanishing peak intensity are not directly connected to the
structural transition from the LTT into the LTO phase, but occur at
a remarkably lower temperature.
\section{Discussion}
Our x-ray results provide direct evidence for charge-stripe order in
$\rm La_{1.45}Nd_{0.4}Sr_{0.15}CuO_4$. Such charge order had previously been
inferred from the magnetic order
observed by neutron scattering \cite{john97}. This is an important result,
because
high-field magnetization \cite{oste97} and muon-spin-rotation ($\mu$SR)
\cite{muon98} studies have indicated that bulk superconductivity also exists in
this sample. Another conclusion of the $\mu$SR study is that static
magnetic order
is present throughout essentially the entire sample volume. Thus, it
appears that
stripe order and superconductivity coexist intimately at $x=0.15$.
Compared to the $x=0.12$ composition, both the charge ordering temperature
and the
superstructure peak intensities are reduced. The superlattice intensities
[normalized to the (200) peak] are roughly 30\%\ weaker in $x=0.15$ relative to
$x=0.12$. A similar reduction in magnetic ordering temperature and
intensity have
been observed previously. At both Sr concentrations, charge ordering
appears at a
higher temperature than magnetic order.
Within the error bars the correlation length of
the stripe order in the $x=0.15$ sample is the same as in the $x=0.12$ sample,
and in both samples the correlation length of the stripes is constant over
a large
temperature range. In contrast to this finding, it is important to note that
neutron scattering and muon-spin rotation experiments revealed that the
spins are
only quasi-static, with a decreasing correlation length, in the temperature
range
above 30~K in the $x=0.12$ sample \cite{muon98,john98}.
This continuous loss of correlation is only observed in
the magnetic signal, whereas
the correlation length for the charge order shows a decrease only very
close to the
disordering temperature. The temperature range in which a residual signal
with an
increased width is observed can be interpreted with fluctuating stripes,
since the detected signal in our x-ray diffraction experiment
is a sum of elastic and inelastic contributions.
One might conclude that fluctuations of stripes set in if the
pinning potential in the LTT phase
is destroyed at the structural transition
from LTT to LTO, as observed in the x~=~0.12 sample,
or if rising temperature competes with a pinning potential via
thermal activation.
In the sample with $x=0.15$, the second possibility holds,
since the suppression of the stripe ordering is not connected
to the structural transition. The pinning potential is smaller in the
$x=0.15$ sample than in the $x=0.12$ sample because
the tilting angle of the oxygen octahedra
is reduced with increasing Sr content in Nd doped LSCO \cite{buch94}.
On the other hand, the idea that the LTT structure is required for
charge-stripe
order is challenged by the interpretation of the Cu NQR results in LSCO by Hunt
{\it et al.} \cite{hunt99}. That work suggests that static stripes can
occur even
within the LTO phase, and that the ordering temperature continues to
increase as
$x$ decreases below 0.12. As an initial test of this picture, we intend to
study
charge order in a Nd-doped crystal with $x=0.10$ in the near future.
To summarize, we have been successful in validating the existence of stripe
order in
$\rm La_{1.45}Nd_{0.4}Sr_{0.15}CuO_4$, a sample which is superconducting
below $\sim 10$~K. Charge-order peaks due to stripe ordering have been observed
at $(2-2\epsilon,0,\pm0.5)$, $(2+2\epsilon,0,0.5)$, and at
$(0,2-2\epsilon,\pm0.5)$. The $2\epsilon$ value of
0.256(1) is in very good agreement with neutron scattering results for the
magnetic peaks. This is a further step in establishing the picture
of static stripes in Nd doped LSCO.
The intensity of the stripe signal decreases almost linearly with
rising temperature and vanishes at 62(5)~K, $\sim8$~K below the structural
transition from the LTT to the LTO phase.
The formation of the stripe pattern in
$\rm La_{1.45}Nd_{0.4}Sr_{0.15}CuO_4$ is less pronounced than in
$\rm La_{1.48}Nd_{0.4}Sr_{0.12}CuO_4$.
\\
Acknowledgements:\\
Work at Brookhaven is supported by
Contract No.\ DE-AC02-98CH10886, Division of Materials Sciences, U.S.
Department of Energy.
|
\section{Introduction}
Numerical studies of QCD on the lattice
provide us a quantitative understanding of the dynamics
of strong interactions.
One of the main goals is to derive the hadron mass spectrum
from the first principle.
In order to give a precise prediction with lattice QCD
simulations we have to control the following
three physical sources of
systematic errors besides the quenching error:
(i) scaling violation, (ii) finite size effects
and (iii) chiral extrapolations.
In the quenched approximation the GF11 collaboration
carried out an extensive
calculation to reduce these three systematic errors and found
that the quenched spectrum is consistent with experiment
within $5-10\%$ errors\cite{GF11}.
The aim of our work is to tell definitely how the quenched
spectrum deviates from experiment, by considerably
diminishing the statistical and systematic uncertainties.
On the other hand,
in spite of much effort devoted in the full QCD spectrum study
we have not yet been able to answer a fundamental question:
to what extent the dynamical quarks affect the light hadron spectrum.
Before embarking upon a realistic full QCD calculation in the future
we attempt to settle this question.
In the first part of this report we present our results
for the quenched light hadron spectrum, examining the validity
of QChPT for the chiral extrapolation.
The two-flavor full QCD results obtained so far are presented
in the second part, where we investigate possible signs of
dynamical quark effects on the light hadron spectrum.
More details are found in Ref\cite{latt98}.
\section{Quenched light hadron spectrum with the Wilson quark action}
\subsection{Details of numerical simulation}
\begin{table}[t]
\caption{\label{tab:para_quench}
Simulation parameters for quenched QCD.}
\begin{tabular}{llllllllllll}
$\beta$ & $L^3\times T$ & $a^{-1}$[GeV] & $La$[fm] & \#conf. &
$\begin{array}{c}{\rm sweep}\\{\rm /conf.}\end{array}$ &
\multicolumn{5}{c}{$m_\pi/m_\rho$} & $\delta$ \\
\tableline
$5.90$ & $32^3\times 56$ & $1.934(16)$ & $3.26(3)$ & $800$ & $200$ &
0.752(1) & 0.692(1) & 0.593(1) & 0.491(2) & 0.415(2) &
0.106(5) \\
$6.10$ & $40^3\times 70$ & $2.540(22)$ & $3.10(3)$ & $600$ & $400$ &
0.751(1) & 0.684(1) & 0.581(2) & 0.474(2) & 0.394(3) &
0.103(6) \\
$6.25$ & $48^3\times 84$ & $3.071(34)$ & $3.08(3)$ & $420$ & $1000$ &
0.760(1) & 0.707(2) & 0.609(2) & 0.502(2) & 0.411(3) &
0.117(7) \\
$6.47$ & $64^3\times 112$ & $3.961(79)$ & $3.18(6)$ & $150$ & $2000$ &
0.759(2) & 0.708(3) & 0.584(3) & 0.493(4) & 0.391(4) &
0.113(13) \\
\end{tabular}
\end{table}
Our simulations are carried out with the plaquette gauge action
and the Wilson quark action.
To control the systematic errors
we carefully choose our run parameters, which are summarized in
Table~\ref{tab:para_quench}. We employ four $\beta$ values
so that the lattice spacing covers the range $a\approx 0.1-0.05$fm
to remove the scaling violation effects
by extrapolation of the data to the continuum limit.
To avoid finite size effects we keep
the physical spatial lattice size approximately constant
at $La\approx 3$fm. A previous study showed that the finite size effects
are $2\%$ or less already at $La\approx 2$fm\cite{finiteV}.
The most subtle issue in controlling the systematic errors
is associated with chiral extrapolations of the hadron masses.
Although the mass formulae predicted by QChPT are considered
to be plausible candidates for the fitting functions
of chiral extrapolations, their validities
should be checked employing a wide range of quark masses.
At each value of $\beta$ we choose five values of the hopping parameter
$K$ corresponding to $m_\pi/m_\rho\approx 0.75$, $0.7$, $0.6$,
$0.5$ and $0.4$. The heaviest two values $m_\pi/m_\rho\approx 0.75$
and $0.7$ are taken to be around the physical strange quark mass.
In terms of these quark masses we calculate hadron masses
both for degenerate and non-degenerate cases.
Gauge configurations are generated with the 5-hit pseudo heat-bath
algorithm incorporating the over-relaxation procedure quadruply
and employing periodic boundary condition.
Quark propagators are solved in the Coulomb gauge both
for the point and the smeared sources with periodic boundary
conditions imposed in all four directions. For the hadron mass
measurement we use the smeared source quark propagators.
The physical point for the degenerate up and down quark mass is
determined by $m_\pi(135.0)$ and $m_\rho(768.4)$.
For the strange quark mass we employ $m_K(497.7)$ or $m_\phi(1019.4)$.
The lattice scale $a^{-1}$ is set with $m_\rho$.
\subsection{Chiral extrapolations}
We first examine the validity of QChPT
investigating the presence of quenched chiral logarithm
in the pseudoscalar(PS) meson sector.
For the quark mass dependence of the PS meson mass
$m_{PS}$ the QChPT predicts\cite{QChPT}
\begin{equation}
\frac{m_{PS}^2}{m_s+m}=A\left\{1-\delta\left[{\rm ln}
\left(\frac{2mA}{\Lambda_\chi^2}\right)
+\frac{m_s}{m_s-m}{\rm ln}\left(\frac{m_s}{m}\right)\right]\right\}
+B(m_s+m)+O((m,m_s)^2),
\label{eq:m_ps}
\end{equation}
where $m$ and $m_s$ are masses of two valence quarks in the PS meson.
This ratio diverges logarithmically
toward the chiral limit due to the $\delta$ term.
To detect the contribution of
the $\delta$ term in a direct manner we introduce
the two variables:
$x=2-(m_s+m){\rm ln}(m_s/m)/(m_s-m)$ and
$y=4 m m_s/(m_s+m)^2\times m_K^4/(m_\pi^2 m_\eta^2)$,
for which eq.(\ref{eq:m_ps}) leads to the relation $y=1+\delta\cdot x$.
Here $\pi$ and $\eta$ are the degenerate PS mesons with
quark mass $m$ or $m_s$ and $K$ is the non-degenerate one with
$m$ and $m_s$. It should be noted that the quark masses are
defined by an extended axial vector current Ward identity(AWI)
$\nabla_\mu A_\mu^{\rm ext}=2m_q^{\rm AWI} P$\cite{m_q^AWI},
where we are free from ambiguities originating from the
determination of the critical hopping parameter $K_c$.
Figure~\ref{fig:delta} shows the distribution of $y$ as a function
of $x$. Our data fall in a wedge shaped with $y=1+0.08 x$
and $y=1+0.12 x$.
A PS meson decay constant ratio $y=f_K^2/(f_\pi f_\eta)$, for which
QChPT predicts $y=1-\delta/2\cdot x$,
is another quantitative test for $\delta$.
Our data indicate that the value of $\delta$ is within
$\delta=0.08-0.16$.
Since we observe clear evidence
for the existence of the quenched chiral
logarithm, we now try to fit the PS meson masses
with the functional form of eq.(\ref{eq:m_ps}).
In Fig.~\ref{fig:ps_chl_quench} we show a typical result
for chiral extrapolations of the degenerate
PS meson mass and the AWI quark mass.
There are two features to be remarked.
One is a good description of the AWI quark mass
with a linear function of $1/K$, which means
the AWI quark mass is proportional
to the vector Ward identity(VWI) quark mass
$m_q^{\rm VWI}=(1/K-1/K_c)/2$.
The other is a good agreement between the critical hopping parameter
$K_c$ determined from the QChPT fit of $m_{PS}^2$
and that from the linear fit of $m_q^{\rm AWI}$.
They are consistent within $2.5\sigma$, which should be compared
with $17\sigma$ ($12\sigma$) discrepancy for the case of
a quadratic (cubic) polynomial fit of $m_{PS}^2$.
Our examinations for the chiral properties of the PS meson masses
strongly suggest the validity of QChPT
with $\delta\approx 0.10(2)$. It would be legitimate to apply
the QChPT mass formulae for the
chiral extrapolation of the vector meson masses
and the baryon ones.
\begin{figure}[b]
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 80mm \epsfbox{kuramashi603fig1.eps}}
\caption[]{\label{fig:delta}
Test of the quenched chiral logarithm using the PS meson masses.}
\vspace*{3.5mm}
\end{minipage}
\hspace{\fill}
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig2.eps}}
\caption[]{\label{fig:ps_chl_quench}
Chiral extrapolations of the degenerate PS meson mass and the
AWI quark mass as a function of $1/K$ at $\beta=5.9$.}
\end{minipage}
\end{figure}
For the vector meson masses we make simultaneous but uncorrelated
fits of the degenerate and non-degenerate data employing
the functional form predicted by QChPT:
\begin{eqnarray}
m_{V}&=&m_V^0+\frac{C_{1/2}}{6}
\left\{\frac{3}{2}(m_\pi+m_\eta)
+2\frac{m_\eta^3-m_\pi^3}{m_\eta^2-m_\pi^2}\right\}
+\frac{C_1}{2}(m_\pi^2+m_\eta^2)
+C_D(m_\pi^3+m_\eta^3)+C_N m_K^3,
\label{eq:m_vfit}
\end{eqnarray}
where $C_{1/2}=-4\pi g_2^2\delta$ with $g_2$ a phenomenological
coupling constant of the vector meson quenched chiral Lagrangian.
It should be noticed that the $O(m_{PS})$ term is
a characteristic of QChPT.
This model function up to the $O(m_{PS}^2)$ terms
describes our vector meson mass data well
as illustrated in Fig.~\ref{fig:v_chl_quench}.
A small bending toward the chiral limit found in the fitting result of
$m_\rho$ reflects the contribution of the $O(m_{PS})$ term.
However, the fitted values for the coefficient $C_{1/2}$
of $O(m_{PS})$ term are much smaller than expected.
The average value of $C_{1/2}$ over four $\beta$ points gives
$C_{1/2}=-0.071(8)$, whose magnitude is ten times smaller than the
phenomenological estimate $C_{1/2}=-4\pi g_2^2\delta\approx -0.71$
with $\delta=0.1$ and $g_2=0.75$.
The magnitude of the $O(m_{PS})$ term in eq.(\ref{eq:m_vfit})
is $0.07\times m_\pi\approx 10$MeV, which means
about a $1\%$ contribution to $m_\rho$.
If we employ the fitting function up to the $O(m_{PS}^3)$ terms,
we find a few times larger value for $C_{1/2}$ at each beta.
In this case, however, the fitting results are very unstable
against the $\chi^2$ value.
\begin{figure}[b]
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig3.eps}}
\caption[]{\label{fig:v_chl_quench}
Chiral extrapolations of vector meson masses
as a function of $m_\pi^2$ at $\beta=5.9$.
Open symbols in the inset represent extrapolated values
at the physical degenerate up and down quark mass.}
\end{minipage}
\hspace{\fill}
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig4a.eps}}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig4b.eps}}
\caption[]{\label{fig:b_chl_quench}
Same as Fig.~\protect{\ref{fig:v_chl_quench}} for octet
and decuplet baryon masses.}
\end{minipage}
\end{figure}
\begin{figure}[b]
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig5a.eps}}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig5b.eps}}
\caption[]{\label{fig:vb_con_quench}
Continuum extrapolations of quenched hadron masses with $m_K$ as
input for the strange quark mass.}
\end{minipage}
\hspace{\fill}
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig6.eps}}
\vskip +.3 cm
\caption[]{\label{fig:spectrum_quench}
Quenched light hadron spectrum in the continuum limit.
GF11 results\protect{\cite{GF11}} are also plotted
for comparison. Horizontal bars denote experimental values.}
\end{minipage}
\end{figure}
For fitting of the octet and decuplet baryon masses
we take the same strategy as for the vector meson mass case.
Assuming the QChPT mass formulae
we fit the degenerate and non-degenerate data
simultaneously without including correlations.
While we find that $O(m_{PS})$ and
$O(m_{PS}^2)$ terms are sufficient to reproduce
our decuplet baryon mass data as shown in Fig.~\ref{fig:b_chl_quench},
for the octet baryon masses a $O(m_{PS}^3)$ term is inevitably required
to explain the convex curvature of the nucleon mass data.
The fitted value for the coefficient $C_{1/2}$ of
the $O(m_{PS})$ term
in the nucleon is $C_{1/2}=-0.118(4)$ as an averaged
value over four $\beta$ points.
This is less than half of the phenomenological estimate
$C_{1/2}=-(3\pi/2)(D-3F)^2\delta\approx -0.27$ with
$\delta=0.1$, $F=0.5$ and $D=0.75$.
The $O(m_{PS})$ term contribution to the nucleon mass
is about $16$MeV or $2\%$.
We find a similar value $C_{1/2}=-0.14(1)$
for $\Delta$.
If the $O(m_{PS}^3)$ terms are included in the fitting function
of the decuplet baryon masses,
we obtain much smaller value for $C_{1/2}$ in $\Delta$.
Whereas we observe that the mass formulae based
on the QChPT reproduce our data adequately,
it is instructive to try another chiral ansatz
employing polynomial fitting functions in $1/K$ (cubic for $N$,
quadratic for others). We employ independent fits
for the different degeneracies in the vector meson and the octet
and decuplet baryons.
The fitting results are shown by dashed lines
in Figs.~\ref{fig:ps_chl_quench}, \ref{fig:v_chl_quench}
and \ref{fig:b_chl_quench}.
Although the QChPT fits seem to give better descriptions
of our hadron mass data than the polynomial ones,
the differences between them are fairly small.
\subsection{Quenched light hadron spectrum in the continuum limit}
The lattice spacing dependences of the hadron masses are
shown in Fig.~\ref{fig:vb_con_quench}. We expect that the leading
scaling violation effects are of $O(a)$ in the Wilson quark case.
A linear extrapolation to the
continuum limit assuming the fitting function $m=m_0(1+\alpha a)$
yields $\alpha \sim 0.2$GeV, for which we find that higher order terms
are safely negligible, {\it e.g.}, $(\alpha a)^2\sim 0.01$
at $a=0.5$GeV$^{-1}$.
While the results of the QChPT fits and the polynomial ones
differ by about $3\%$($5\sigma$) in the largest case
after the chiral extrapolation at each $\beta$,
the differences in the continuum limit are
within $1.5\%$($1.5\sigma$) of the results of the QChPT fits.
At a few percent level of statistical errors it is hard to appreciate
the differences between the QChPT chiral extrapolations
and the polynomial ones.
Our final result of the quenched light hadron spectrum is presented
in Fig.~\ref{fig:spectrum_quench}. The spectrum deviates from experiment
systematically and unambiguously.
The following discrepancies should be noticed:
The $K$-$K^*$ meson hyperfine
splitting is definitely underestimated by $9.5\%$($4.3\sigma$) for
the $m_K$ input case and by $16\%$($6.1\sigma$) for the $m_\phi$ input case;
The nucleon mass is appreciably smaller than experiment
by $7\%$($2.5\sigma$);
For the strange octet baryon masses our data
with the $m_K$ input are systematically
smaller than experiment by $6-9\%$($4-7\sigma$),
which is much reduced in the $m_\phi$ input case;
While the $m_K$ input leads $30\%$ smaller estimates for
the decuplet mass splittings in average, we observe rather
good agreement for the $m_\phi$ input results.
\section{Light hadron spectrum in two-flavor full QCD}
\subsection{Details of numerical simulation}
\nopagebreak
The simulation of full QCD requires a huge amount of
computing time compared to the quenched approximation.
In order to secure an adequate physical spatial lattice size
for avoiding finite size effects, the lattice spacing
is compelled to be coarse $a^{-1}\rlap{\lower 3.5 pt\hbox{$\mathchar \sim$}}\raise 1pt \hbox {$<$} 2$GeV, which
urges us to employ improved actions.
Our choice is a RG-improved gauge action\cite{RGaction}
and the SW quark action\cite{SWaction} with a mean-field improved value
of $c_{SW}$. This decision is based on a former comparative
study of various combinations of improved gauge and quark
actions at $a^{-1}\approx 1$GeV\cite{prestudy}.
Our full QCD study is performed with two flavors of sea quarks
which are supposed to be the degenerate dynamical up and down quarks.
Our simulation parameters are summarized in Table~\ref{tab:para_full}.
Four $\beta$ values covering the lattice spacing
in the range $a^{-1}\approx 1-2$GeV are employed to
examine the scaling behavior of the light hadron spectrum.
We keep the physical spatial lattice size approximately constant
at $La\approx 2.4$fm except for the finest lattice where
$La\approx 2.0$fm.
At each value of $\beta$ we choose four values of sea quark mass
corresponding to $m_\pi/m_\rho\approx 0.8$, $0.75$, $0.7$ and $0.6$.
Gauge configurations including the dynamical sea quark effects
are generated with the HMC algorithm.
We calculate hadron masses
using the five valence quarks whose masses correspond to
$m_\pi/m_\rho\approx 0.8$, $0.75$, $0.7$, $0.6$ and $0.5$.
Hadron propagators are constructed with
the smeared source quark propagators
at every fifth trajectory. Hadron masses are extracted by
an exponential fit ignoring correlations between time slices.
Errors are estimated by the jackknife
method with a bin size of 10 configurations (50 HMC trajectories).
The light hadron spectrum is obtained by setting the
sea quark mass on the physical up and down quark mass
and the valence quark mass on the physical up
and down quark mass or the strange one.
The physical point of the degenerate up
and down quark mass is fixed with $m_\pi$ and $m_\rho$ as input.
For the strange quark mass the physical point is determined from
$m_K$ or $m_\phi$.
The lattice scale $a^{-1}$ is set with $m_\rho$.
\subsection{Chiral extrapolations in the sea and valence quark masses}
In Fig.~\ref{fig:msn_chl_full} we plot a result
for the sea and valence quark mass dependences of
hadron masses at $\beta=1.95$. The data are parameterized
by the averaged hopping parameter
$1/K_{av}=(1/K_{val(1)}+1/K_{val(2)})/2$ of the two kind of
valence quarks constituting the hadrons.
$S$ represents a valence quark with $K_{val}=K_{sea}$ and
$V$ a valence quark with $K_{val}\ne K_{sea}$.
\begin{table}[t]
\caption{\label{tab:para_full}
Simulation parameters for full QCD.}
\begin{tabular}{lccccllll}
$\beta$ & $L^3\times T$ & $c_{SW}$ & $a$[fm] & $La$[fm]
& \multicolumn{4}{c}{$m_{\pi}/m_{\rho}$ for sea quarks} \\
& & & &
& \multicolumn{4}{c}{\#traj.} \\
\tableline
1.80 & $12^3\times 24$ & $1.60$ & 0.215(2) & 2.58(3)
& 0.8060(7) & 0.753(1) & 0.696(2) & 0.548(4) \\
&&&&& 6250 & 5000 & 7000 & 5250 \\
1.95 & $16^3\times 32$ & $1.53$ & 0.153(2) & 2.45(3)
& 0.8048(9) & 0.751(1) & 0.688(1) & 0.586(3) \\
&&&&& 7000 & 7000 & 7000 & 5000 \\
2.10 & $24^3\times 48$ & $1.47$ & 0.108(2) & 2.59(5)
& 0.806(2) & 0.757(2) & 0.690(3) & 0.575(6) \\
&&&&& 2000 & 2000 & 2000 & 2000 \\
2.20 & $24^3\times 48$ & $1.44$ & 0.086(3) & 2.06(6)
& 0.800(2) & 0.754(2) & 0.704(3) & 0.629(5) \\
&&&&& 2000 & 2000 & 2000 & 2000 \\
\end{tabular}
\end{table}
We observe that the data of the meson masses
are distributed on different four lines
in accordance with the four sea quarks.
Partially quenched data on each sea quark
show almost linear behavior both for $m_{PS}^2$ and $m_{Vec}$.
Their slopes, however, slightly depend on the
sea quark masses: As the sea quark mass decreases
the slope decreases for $m_{PS}^2$,
while the slope increases for $m_{Vec}$.
In the $S$-$S$ channel
$m_{PS}^2$ behaves almost linearly in
$1/K_{sea}$, whereas $m_{Vec}$ exhibit a convex curvature.
Chiral extrapolations of the meson masses
are made by global fits of all the data
assuming quadratic functions
in terms of $1/K_{sea}$ and $1/K_{val}$.
For the PS meson mass we employ
\begin{eqnarray}
m_{PS}^2&=& B_s m_{sea}+B_v{\overline m}_{val}+C_s m_{sea}^2
+C_v{\overline m}_{val}^2+C_{sv}m_{sea}{\overline m}_{val}
+C_{12}m_{val(1)}m_{val(2)},
\end{eqnarray}
where bare quark masses are defined by
$m_{sea/val(i)}=(1/K_{sea/val(i)}-1/K_c)/2$ and
${\overline m}_{val}$ is the averaged value of two
valence quark masses.
Similar fitting function without the valence-valence cross term
$m_{val(1)}m_{val(2)}$ is used for
the vector meson masses and the decuplet baryon masses.
The fitting results for the meson masses are drawn
in Fig.~\ref{fig:msn_chl_full}.
As for the octet baryon masses we find a rather complicated
situation in Fig.~\ref{fig:msn_chl_full}: Partially quenched
baryon masses are not functions of ${\overline m}_{val}$.
For the fitting function we take a combination of linear terms
based on the ChPT prediction and general quadratic terms of
individual $m_{val(i)}$. This function has 12 free parameters
in all, which are determined by a combined
fit of $\Sigma$-like and $\Lambda$-like baryon
masses.
\begin{figure}[b]
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 68.5mm \epsfbox{kuramashi603fig7a.eps}}
\centerline{\epsfxsize 68.5mm \epsfbox{kuramashi603fig7b.eps}}
\centerline{\epsfxsize 68.5mm \epsfbox{kuramashi603fig7c.eps}}
\caption[]{\label{fig:msn_chl_full}
Sea and valence quark mass dependence of hadron masses
at $\beta=1.95$. Solid and broken lines represent
results of quadratic fits for chiral extrapolations.
See text for the labels $S$ and $V$.}
\end{minipage}
\hspace{\fill}
\begin{minipage}[b]{80mm}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig8a.eps}}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig8b.eps}}
\vspace*{1mm}
\centerline{\epsfxsize 73mm \epsfbox{kuramashi603fig8c.eps}}
\caption[]{\label{fig:spectrum_full}
Light hadron spectrum in two-flavor full QCD
with $m_K$ as input for the strange quark mass
as a function of the lattice spacing.
For comparison quenched QCD results in the last section
are also plotted.}
\end{minipage}
\end{figure}
\subsection{Sea quark effects on the light hadron spectrum}
The results for the light hadron spectrum are presented
in Fig.~\ref{fig:spectrum_full}
as a function of the lattice spacing. For comparative purpose
we also plot the quenched results(open symbols)
from the last section, which would help us distinguish
the dynamical sea quark effects.
We observe an intriguing feature in the meson sector:
The $K^*$ and $\phi$ meson masses in full QCD are heavier than
those in quenched QCD at finite lattice spacing and become closer
to experimental values as the lattice spacing decreases.
However, our full QCD results do not necessarily
agree with experiment in the continuum limit.
We have two possible reasons.
One is the relatively heavy sea quark masses corresponding
to $m_\pi/m_\rho\approx 0.6-0.8$, for which
the chiral extrapolations are rather ambiguous.
The other is the quenched treatment of the strange quark.
For the baryon sector it is hard to read any meaningful
implication in comparison between the full QCD results
and the quenched ones. At the finite lattice spacing
the full QCD spectrum is consistent with the quenched ones
within rather large errors, while the lattice spacing dependence
of the decuplet baryon masses in full QCD
is opposite to that in quenched case.
\section{Conclusions}
We have presented our results for the quenched
light hadron spectrum. After examining the validity of
QChPT in the pseudoscalar meson masses
we applied its hadron mass formulae
to the chiral extrapolations.
Although the QChPT predictions reproduce
our hadron mass data well,
for the vector meson and baryon masses
it is hard to confirm the contributions of the term
specific to QChPT.
The spectrum in the continuum limit deviates
systematically and unambiguously from experiment
on a level of about $10\%$ far beyond the statistical error
of $1-2\%$ for mesons and $2-3\%$ for baryons.
Our full QCD study is an exploring step
toward a realistic QCD simulation.
For the light hadron spectrum we found an
encouraging result that the deficiency of
the meson hyperfine splitting in the quenched approximation
is largely compensated with the dynamical sea quark effects.
At this stage, however, the most important conclusion
is as follows:
To perform a close investigation of the sea quark effects
we need a direct comparison between the full QCD results and
the quenched ones employing the same gauge and quark actions
with the same $a^{-1}$, $m_\pi/m_\rho$ and $La$.
Work in this direction is now in progress.
I am grateful to all the members of the CP-PACS Collaboration for
their help in preparing this manuscript.
I would particularly like to thank R.~Burkhalter and T.~Yoshi{\'e}.
This work is supported in part by the Grants-in-Aid of Ministry
of Education(Nos. 08NP0101 and 09304029).
|
\section{Introduction}
In spite of the extraordinary successes of the Standard Model (SM)
in explaining all the high energy experiments with large luminosity
up to now \cite{Altarelli98},
there have been various attempts to search for physics beyond the SM.
Even though some astrophysical and collider measurements
have intimated that the SM is not the whole story,
the signals have been regarded as hints, not as facts.
A major portion of the motivations to extend the SM comes
from conceptual discomfort in the theoretical viewpoints,
such as the hierarchy problem:
the SM cannot provide a satisfactory answer to why the nature
allows such an enormous ratio between two fundamental mass scales,
the electroweak scale at $\sim 100$ GeV and the Planck mass scale
at $\sim 10^{19}$ GeV.
This problem has prompted the extensive studies of the supersymmetric
models \cite{Susy}, the technicolor models \cite{Technicolor}, etc.
Recently Arkani-Hamed, Dimopoulos, and Dvali (ADD)
have approached the hierarchy problem
in a salient way by removing one prerequisite of the problem itself:
the Planck mass is not fundamental, that is,
there exists only one fundamental mass scale, $M_S$, in the nature \cite{ADD1}.
The observed extremely small Newton's constant $G_N$
or alternatively huge Planck mass scale can be obtained by
introducing the existence of extra compact $N$ dimensions.
Then the Planck mass is related to the $M_S$ and the size of
extra dimensions.
For example, if the extra dimensional space is the $N$-dimensional
torus with the same compactification radii $R$, the relation is
\begin{equation}
\label{Planck}
\kappa^2 R^N = 16 \pi (4 \pi)^{N \over 2} \Gamma\left(
\frac{N}{2}
\right)
M_S^{-(N+2)}
\,,
\end{equation}
where $\kappa^2\equiv 16 \pi G_N$ \cite{Han}.
The excellence of Newtonian mechanics in describing the solar system excludes
the $N=1$ case where the $R$ is order of $10^{13}$ cm.
The $N=2$ case which implies mm scale extra dimensions
is not excluded by the current macroscopic measurement of
gravitational force \cite{macro}.
The cases of $N>2$ are also acceptable but difficult to probe through
macroscopic observations.
In order to be phenomenologically consistent with
the SM at the electroweak scale, the model
assumes that the SM fields are confined in our 4-dimensional
brane while gravitons are freely propagating in the whole $(4+N)$-dimensional
bulk.
Even though this discrimination can be achieved by considering
our world as a topological defect of a higher dimensional theory
and by localizing the SM fields at the vortex \cite{ADD1},
its natural realization had been already
discussed in string theories \cite{ADD2,Sundrum}.
Strong string coupling obtained by T-dualising
transforms the Kaluza-Klein modes of opens strings into
the winding modes of open strings of which the two ends are attached on
a brane, which are good candidates of the SM particles.
And closed strings are still able to propagate orthogonal to the brane,
identified as gravitons.
The existence of extra dimensions reveals through the interactions
between gravitons and the SM particles on our brane.
Although the required invariance under general coordinate
transformations in the brane and bulk
can, in principle, specify the interactions,
the presence of non-trivial metric in the extra dimension
complicates them.
Unless gravitons take momentum compatible with or larger than $M_S$
at high energy collisions,
the spacetime region where collisions occur can be regarded as flat
\cite{Wells}.
Then the linear approximation is valid so that
the metric in $(4+N)$-dimensions can be expanded around the Minkowski
metric as
\begin{equation}
\label{2}
\hgmn = \hetamn +\hkp \hhmn
\,,
\end{equation}
where the $\hhmn$ is the canonical graviton fields and the hatted indices
denote the $(4+N)$-dimensional spacetime.
After the compactification on a sub-manifold of extra dimensions,
our brane effectively possesses Kaluza-Klein towers of
massive spin 2 gravitons $\thmnn$,
massive vector particles $\tilde{A}_{\mu i}^\vn $,
and massive spin zero particles $\tilde{\phi}_{ij}^\vn $.
The matter-graviton couplings are specified by the minimal coupling of gravity,
yielding the Feynman rules of the interactions
between the Kaluza-Klein gravitons and the SM
particles
\cite{Sundrum,Han,Wells}.
Following the results of Ref.\cite{Han},
we use the effective action to the leading order in $\kappa=1/M_{\rm pl}$
\begin{equation}
\label{3}
I=-\frac{\kappa}{2}
\sum_\vn
\int d^4 x
\left[
\tilde{h}^{\mu\nu,\vn} T_{\mu\nu} + w \tilde{\phi}^\vn T^\mu_\mu
\right]
\,,
\end{equation}
where $w=\sqrt{2/3(n+2)}$ and $T^{\mu\nu}$ is the energy-momentum tensor.
It is to be noted that new kinds of interactions due to the low scale
quantum gravity are neutral current ones.
Recently the idea of the existence of
large extra dimensions has drawn quite explosive attentions of
particle physicists.
Above all, attractive is that the quantum gravity scale can be as low
as TeV so to be testable.
It would be worthwhile to search for collider tests
which can confirm or exclude the validity of the model.
Various phenomenological studies have been performed to
constrain the scale of $M_S$ or the number of extra dimensions $N$
from the existing data of collider experiment \cite{Peskin,Hewett,Rizzo}, and
in the future experiments \cite{Hewett,Rizzo,top,Agashe,Balazs,Cheung}.
Two kinds of processes at colliders are computable by using the
effective action in Eq.(\ref{3}),
single graviton emission and virtual exchange
of gravitons.
For the first case, the presence of extra dimension
induces the Kaluza-Klein multiplicity of phase factor,
proportional to $(\Delta E \cdot R)^N$,
where $\Delta E$ is the energy of emitting graviton.
Thus any process involving single graviton emission
shows the large dependence on $N$ since the branching ratio is proportional to
$(\Delta E/M_S)^{N+2}$ \cite{ADD1}.
Studies of various processes
show that even in the future NLC or LHC,
the $N\geq 5$ cases are practically impossible to
distinguish from the SM background \cite{Hewett,Rizzo,top,Agashe}.
Virtual exchange of gravitons does not have
such dependence;
the scattering amplitudes,
suppressed by $\kappa^2$ and mediated by Kaluza-Klein towers,
are proportional to $1/M_S^4$.
Their dependences on the $N$, according to compactification models,
may not be as critical as in the single graviton emission case
\cite{Hewett,Rizzo}.
Moreover in this case,
the polarizations of incoming and outgoing particles are affected by
the fact that gravitons are of spin two while photons or $Z$ bosons,
mediating the SM neutral currents, are of spin one.
Therefore, one of the most useful and significant measurements to
probe the existence of the extra dimensions would be the measurements of
the polarizations at virtual graviton exchange processes.
In this paper we concentrate on the process $e^+ e^- \to W^+ W^-$
in the future Linear Colliders (LC) \cite{LC}.
The spin-one nature of the $W$ bosons provides
more channels to signal new physics, and the measurements
of the $W$ polarizations are attainable
from the decay angular distributions \cite{Wpol}.
Therefore it is desirable to derive in detail the effects
of large extra dimensions on the $W$ and beam polarizations.
\vskip 1.5cm
\begin{center}
\begin{picture}(230,100)(0,0)
\Text(55, 5)[]{(a)}
\Text(15,100)[]{$e^-$}
\Text(15,20)[]{$e^+$}
\Text(40,60)[]{$\nu_e$}
\ArrowLine(10,90)(50,90)
\ArrowLine(50,90)(50,30)
\ArrowLine(50,30)(10,30)
\Photon(50,90)(90,90){3}{7}
\Photon(50,30)(90,30){3}{7}
\Text(95,100)[]{$W^-$}
\Text(95,20)[]{$W^+$}
\Text(170, 5)[]{(b)}
\ArrowLine(120,90)(150,60)
\ArrowLine(150,60)(120,30)
\Text(125,100)[]{$e^-$}
\Text(125,20)[]{$e^+$}
\Text(170,80)[]{$\gamma,Z,\tilde{h}_{\mu\nu}^{\vec{n}}$}
\Photon(150,60)(190,60){3}{5}
\Photon(190,60)(220,90){3}{5}
\Photon(190,60)(220,30){3}{5}
\Text(215,100)[]{$W^-$}
\Text(215,20)[]{$W^+$}
\end{picture}
\end{center}
\smallskip
{ }
\noindent
Figure~1: {\it Feynman Diagrams contributing to
the process $e^+e^-\rightarrow W^+ W^-$ including
large extra dimension effects.
}
\smallskip
\smallskip
For the process
\begin{equation}
e^-(p_1,\kp)+e^+(p_2,\overline{\kp})\to
W^-(q_1,\lm)+W^+(q_2,\overline{\lm})
\,,
\end{equation}
there are four Feynman diagrams, one $t$-channel diagram mediated
by the neutrino and three $s$-channel ones mediated by
the photon, $Z$ boson, and spin-2 gravitons as depicted in Fig.1.
At high energies where the electron mass is negligible,
there are 18 different helicity amplitudes
${\mathcal M}^\pm (\lm,\overline{\lm})$,
where superscripts denote the electron helicities.
Since the CP-invariance relates some amplitudes such as
\begin{equation}
{\mathcal M}^\pm (\lm,\overline{\lm})
= {\mathcal M}^\pm (-\overline{\lm},-\lm)
\,,
\end{equation}
thus only 12 amplitudes are independent \cite{Denner}.
By using the helicity formalism in Ref.~\cite{HAmp} and
the polarization convention in Ref.~\cite{Denner},
we calculate the helicity amplitudes for the left-handed
electron beam as
\begin{eqnarray}
\label{LH}
{\mathcal M}^-_{++} &=&
s_\theta [
f_L\bw + \frac{g^2 s}{4 t}(c_\theta-\bw)
-2f_D c_\theta(1-\bw^2) ]
\,, \\ \no
{\mathcal M}^-_{+0} &=&
-\frac{\sqrt{2}(1+c_\theta)}{\mw}[
f_L\bw + \frac{g^2 s}{8 t}
(2c_\theta - 1-2\bw+\bw^2)
-f_D(2c_\theta-1)(1-\bw^2)]
\,, \\ \no
{\mathcal M}^-_{+-} &=&
2 s_\theta (1+c_\theta)[
\frac{g^2 s}{8 t} -f_D
]
\,, \\ \no
{\mathcal M}^-_{0+} &=&
\frac{\sqrt{2}(1-c_\theta)}{\mw}[
f_L\bw + \frac{g^2 s}{8 t}
(2c_\theta+1-2\bw-\bw^2)
-f_D(2c_\theta+1)(1-\bw^2)]
\,, \\ \no
{\mathcal M}^-_{00} &=&
-\frac{s_\theta}{\mw^2}
[
f_L\bw (3-\bw^2) + \frac{g^2 s}{4 t}
(2c_\theta-3\bw+\bw^3)
-2f_D c_\theta(2-3\bw^2+\bw^4)
]
\,, \\ \no
{\mathcal M}^-_{-+} &=&
-2s_\theta (1-c_\theta)
[
\frac{g^2 s}{8 t} -f_D
]\,,
\end{eqnarray}
and for the right-handed electron beam we have
\begin{eqnarray}
\label{RH}
{\mathcal M}^+_{++} &=&
s_\theta[
f_R\bw - 2f_D c_\theta(1-\bw^2)
]
\,, \\ \no
{\mathcal M}^+_{+0} &=&
\frac{\sqrt{2}(1-c_\theta)}{\mw}[
f_R\bw
-f_D(2c_\theta+1)(1-\bw^2)
]
\,, \\ \no
{\mathcal M}^+_{+-} &=&
2 f_D s_\theta(1-c_\theta)
\,, \\ \no
{\mathcal M}^+_{0+} &=&
-\frac{\sqrt{2}(1+c_\theta)}{\mw}[
f_R\bw
-f_D(2c_\theta-1)(1-\bw^2)
]
\,, \\ \no
{\mathcal M}^+_{00} &=&
-\frac{s_\theta}{\mw^2}[
f_R\bw(3-\bw^2)-2f_D c_\theta(2-3\bw^2+\bw^4)
]
\,, \\ \no
{\mathcal M}^+_{-+} &=&
-2 f_D s_\theta(1+c_\theta)\,.
\end{eqnarray}
The CP-invariance implies
\begin{equation}
{\mathcal M}^\pm_{--}={\mathcal M}^\pm_{++}, \quad
{\mathcal M}^\pm_{-0}={\mathcal M}^\pm_{0+}, \quad
{\mathcal M}^\pm_{0-}={\mathcal M}^\pm_{+0}.
\end{equation}
In Eqs.(\ref{LH}) and (\ref{RH})
we denote $s_\theta=\sin\theta$, $c_\theta=\cos\theta$, and
\begin{eqnarray}
\es_L&=& -\hf +\sin^2\theta_W,\quad
\es_R =\sin^2\theta_W\,,
\\ \no
f_{L,R} &=& \frac{g^2 \es_{L,R}}{1-M_Z^2/s} -e^2\,,\quad
\\ \no
\mw &=& \frac{M_W}{\sqrt{s}/2} =\sqrt{1-\beta^2_W} \,.
\end{eqnarray}
The terms proportional to $f_{L,R}$ denote the contributions
from the $\gamma$- and $Z$-mediated diagrams,
and those proportional to the $g^2 s/t$
from the neutrino-mediated one.
It can be easily seen that the preparation of
right-handed electron beam switches off
the $t$-channel $\nu$-mediated contributions.
These SM results are consistent with those in Ref.\cite{Denner,Agashe}.
The low scale quantum gravity effects are included in $f_D$, defined by
\begin{eqnarray}
\label{f_D}
f_D \equiv -\frac{\pi s^2}{2 M_{\rm pl}^2 } \sum_{\vn}
\frac{1}{s-m_\vn^2}
&\simeq &
\frac{\pi s^2}{2M_S^4} \ln (
\frac{M_S^2}{s}
)
~\quad {\rm for} ~~ N=2 \\ \no & \simeq&
\frac{\pi s^2}{(N-2) M_S^4} \quad\quad {\rm for} ~~ N>2
\,.
\end{eqnarray}
It is to be noted that the two helicity amplitudes
${\mathcal M}^+_{+-} $ and
${\mathcal M}^+_{-+} $ vanish at the tree level in the SM,
but retain extra dimension effects as sizable as
the other amplitudes.
In Fig.~2 and 3, we plot the differential cross sections for the
$W$ pair production at $e^+ e^-$ collisions
with respect to the $W^-$ scattering
angle against the electron beam at $\sqrt{s}=1$ TeV
in the case of $N=2$ and $M_S=2.5$ TeV, broken down to the transverse and
longitudinal helicity components of the $W$ bosons.
The cases with $N>2$ unless $N$ is too large
shows similar behaviors.
The effects of large extra dimensions with respect to the SM background
can be enhanced by using the right-handed
electron beam and selecting the $W^+ W^-$ polarizations
to be both transverse.
This is because the employment of the right-handed electron beam eliminates
the dominant SM contributions of
the $t$-channel $\nu$-mediated diagram in Fig.1.
And the SM background from the $s$-channel
with the right-handed electron beam is proportional to
\begin{equation}
f_R =e^2 \left[
\frac{1}{1-M_Z^2/s}-1
\right]
\propto \frac{M_Z^2}{s}\qquad{\rm for}\quad s \gg m_Z,
\end{equation}
which $decreases$ as the beam energy becomes larger,
while the extra dimension correction proportional to $f_D$ in Eq.(\ref{f_D})
$increases$.
Furthermore, at the tree-level,
${\mathcal M}^{+{\rm (SM)}}_{+-}={\mathcal M}^{+{\rm (SM)}}_{-+}=0$
in the SM.
The $\sigma_{_{\rm TT}}$ including large extra dimension effects
is about $\sim10^4$ times the SM background.
In practice,
the generation of 100\% polarized electron beam is infeasible.
We consider the expected polarization of the electron beam as 90\%.
Figure 4 shows the differential cross sections
against the $W$ scattering angle
according to the $W$ polarizations at $\sqrt{s}=1$ TeV when
$N=2$ and $M_S=2.5$ TeV.
The dominant $t$-channel SM background contaminates the unique behavior of
$d \sigma_{TT}$, however, substantial corrections to the SM background
still remain.
We do not consider the polarization of the positron beam
since it is more difficult to generate, expected presumably
in the range of 60 $\%$ to $65\%$ \cite{LC}.
No radiative corrections are included here.
The SM radiative corrections have the effects of the same orders
of magnitude on the results with or without considering this model
since the effects of the extra dimensions
mainly come from the interference with the SM amplitudes.
\vskip 1.0cm
\noindent
\begin{center}
\begin{tabular}{|c|cc|cc|}\hline
&\multicolumn{2}{|c|}{~~$\sqrt{s}=0.5$ TeV
( $\int{\mathcal L} = $50 fb$^{-1}$)~~}
&\multicolumn{2}{|c|}{~~$\sqrt{s}=1$ TeV
( $\int{\mathcal L} = $200 fb$^{-1}$)~~}
\\ \hline
{} &~~ $N=2$~~ &~~ $N=6$~~ & ~~$N=2$~~ &~~ $N=6$~~ \\ \hline
~~$\sigma_{\rm tot}^{\rm unpol}$ ~~& 4.6 & 2.6 & 9.3 & 5.3 \\
$\sigma_{\rm TT}^{\rm unpol}$ & 4.6 & 2.6 & 9.3 & 5.4 \\
$~~\sigma_{\rm tot}^{\rm 90\%\,RH}~~$ & 3.4 & 2.0 & 6.8 & 4.1 \\
$\sigma_{\rm TT}^{\rm 90\%\,RH}$ & 3.4 & 2.0 & 7.0 & 4.1 \\
$\sigma_{\rm tot}^{\rm 90\%\,LH}$ & 5.1 & 2.9 & 10.1 & 5.8 \\
$\sigma_{\rm TT}^{\rm 90\%\,LH}$ & 5.1 & 2.9 & 10.2 & 5.8 \\
\hline
\end{tabular}
\end{center}
{Table~1}. {\it The LC bounds of $M_S$ in TeV at 95\%
confidence level according to the beam or $W$ pair polarizations.}
\vskip 0.5cm
The LC bounds on the $M_S$ are derived by the statistical errors
with the angular cut $|\cos\theta|<0.95$ at 95\% confidence level
at $\sqrt{s}=0.5$, 1.0 TeV and $N=2,6$
from six different observables according to
the beam and $W$ pair polarizations.
As for the beam polarization effects on the $M_S$ bounds,
the preparation of the left-handed electron beam is expected to yield
higher bounds.
With a given beam polarization,
the $\sigma_{_{\rm TT}}$'s are likely to give higher $M_S$ bounds.
On account of the smaller numbers of transversely polarized $W$ pair events,
these results imply that large extra dimension corrections
in the $\sigma_{_{\rm TT}}$'s are much larger than those in the
$\sigma_{\rm tot}$.
It is concluded that valuable information about the models with
large extra dimensions can be obtained
by observing the $W$ pair and beam polarizations
at the $e^+ e^-\to W^+ W^-$ process.
In particular,
the measurements of the cross section for transversely polarized $W$ pair
with the right-handed electron beam highly enhances the possibilities
to see the low scale quantum gravity effects.
The current inability to generate purely polarized beam at $e^+ e^-$ colliders
contaminates this feature.
Almost purely polarized beams are possible in the future $\mu^+ \mu^-$
colliders \cite{muon-collider},
since the muons are prepared through the pion decays
accompanied by purely chiral neutrinos.
We expect definite signal of the large extra dimension effects through
the observations of transverse polarizations of $W$ pair
with the right-handed muon beam at the muon colliders.
It has been shown that for the LC bounds of the string scale $M_S$
the use of left-handed electron beam is preferred,
and for the probe of large extra dimension effects
the measurements of the cross section for transversely polarized $W$ pair are.
\acknowledgments
We would like to appreciate valuable discussions with S.Y. Choi
and T. Lee.
The work was supported in part by the Korea Science
and Engineering Foundation (KOSEF) through the Center for Theoretical Physics,
Seoul National University.
HSS would like to acknowledge the financial support of the
Korea Research Foundation through the 97 Sughak Program and
98--015-D00054.
\bigskip
\bigskip
|
\section*{ ACKNOWLEDGMENTS }
\end{center}
The authors would like to thank the partial financial support by
National Natural Science Foundation of China under Grant Nos.
19675044 and 19875026.
|
\section{Introduction}
The Weyl representation places quantum mechanics in phase space. The density
operator is mapped into a real phase function that projects onto the
position or momentum probability densities. The unitary operators
corresponding to linear canonical transformations transform the Weyl symbol
of any operator as a classical variable. Therefore, it is not surprising
that the study of the semiclassical limit of nonintegrable systems has
relied heavily on the Weyl representation as reviewed by reference \cite
{ozrep}.
The study of systems that are chaotic in the classical limit has developed
several models supported by a compact phase space. The simplest choice is
that of a $\left( 2L\right) $-dimensional torus, corresponding to a system
with $L$ degrees of freedom. Indeed, if the system propagates in discrete
time ( a mapping of the torus onto itself) even the special case $L=1$ may
be chaotic as is the case of the cat map \cite{arnold}, \cite{hanay} or the
baker's map \cite{balavoros}, \cite{saraceno}. It is well known that the
Hilbert space corresponding to a classical torus is finite. We may picture
the $N$ allowed positions and momenta as forming a lattice, the {\it quantum
phase space} (QPS), even though no rigorous definition of position and
momentum operators is available on the torus \cite{pqtor}. The semiclassical
limit is then obtained as $N\sim \left( 2\pi \hbar \right) ^{-1}\rightarrow
\infty .$
Though it is a great advantage to investigate numerically the propagation of
finite vectors or matrices defined in the QPS, it is somewhat disconcerting
that the difference between classical and quantum motion is more extreme on
the torus than on the plane. This problem also manifests itself in the
adaptation of the Weyl representation to the torus. The existing literature
\cite{wooters}-\cite{kaperpeev} relies mainly on formal procedures, so that
the achievement of the classical limit as $N\rightarrow \infty $ may not be
preceded by the emergence of classical structures even for finite $\hbar $,
such as have been found for a plane phase space.
Clearly, a way to avoid this difficulty is to consider the classical torus
as a periodic plane phase space, to quantize the latter and then to project
this on QPS. We thus generalize the procedure of Hannay and Berry \cite
{hanay}, allowing for arbitrary ''Bloch '' or ''Floquet ''angles for each
circuit of the torus. It is then possible to project the appropriate plane
translation and reflection operators onto the torus. Hence, we define the
Weyl (or center) representation and its conjugate chord representation while
maintaining the main geometrical features characteristic of the plane.
In section 2 we present the translation operators of momentum and position
for a torus considered as the fundamental unit cell of the periodic plane.
By also allowing tori made up of more than one cell, we show that the unit
quantum torus may be obtained as the projection of the Hilbert space
corresponding to such a larger torus. Finally, by connecting the Hilbert
space for the plane to that of an infinitely large torus, we obtain the
projection of plane operators onto the torus.
Section 3 summarizes important features of the Weyl representation on the
plane. Defining the translation operators and their Fourier transform, the
reflection operators, we derive the Weyl symbols and their product rules in
terms of integrals over phase space polygons.
In section 4 we project appropriate translation operators onto QPS. However,
it is found that the reflection operators are supported on a lattice with
half the spacing of QPS . In consequence, the trace of these operators is
not homogeneous, which leads to complications in the formulae for products
of operators. Only in the case that $N$ is odd, can we simplify the product
rule for the Weyl symbols into a form that is analogous to the theory in the
plane. In any case, we need only half the number of sums derived in the
previous work of Galleti and Toledo Pisa \cite{galeti1} and these sums
depend on the symplectic areas of the same polygons arising in the plane
theory. Finally, we discuss the restricted form of symplectic invariance
that holds for the Weyl and chord representation on the torus: The Weyl
symbols transform classically under the action of quantum cat maps.
Though we have motivated our paper through discrete time models, periodic
Hamiltonians have obvious applications in solid state physics. Thus, section
5 is dedicated to the derivation of a path integral for the Weyl symbol of
the propagator. This relies on the symplectic area of polygons as in the
plane theory. In the semiclassical limit, the propagator is expressed in
terms of the center generating function presented in reference \cite{ozrep}.
Throughout this work we differentiate operators on the plane by italic $%
\widehat{A}$, as opposed to bold operators $\widehat{{\bf A}}$ on the torus.
\section{Hilbert space for tori}
\setcounter{equation}{0}
Classical phase space of ($2L)$ dimensions may be considered to be periodic,
so that we confer to it the topology of a $(2L)$ dimensional torus.
Evidently, we may use invariance, with respect to symplectic
transformations, to equate the periods $\Delta q=\Delta p=\nu $ of the
position and momentum coordinates. The usual choice is $\nu =1$, but we will
leave this as a free integer parameter so as to study the nesting of tori,
that is, the case where quantization is imposed on a larger (periodic)\
region than the unit cell. Thus, the number of unit cells will be $\nu
^{2L}. $
It is important to treat the specification of the Hilbert space of quantum
states for the torus, or {\it prequantization}, independently from the
dynamics of the system. That is, we treat the quantum kinematics,
corresponding to the geometrical description of phase space at the classical
level. A complete description for prequantization must include boundary
conditions; which are here that the wave functions satisfy Bloch conditions:
\begin{eqnarray}
{\bf \Psi }(q+\nu ) &=&e^{2\pi i\chi _p}{\bf \Psi }(q), \label{eq1} \\
{\bf \tilde{\Psi}}(p+\nu ) &=&e^{-2\pi i\chi _q}{\bf \tilde{\Psi}}(p)
\label{eq:eq2}
\end{eqnarray}
where
\begin{equation}
{\bf \tilde{\Psi}}(p)=(2\pi \hbar )^{-1/2}\int e^{-ipq/\hbar }{\bf \Psi }%
(q)dq \label{eq:91}
\end{equation}
and $2\pi i\chi _p$ and $2\pi i\chi _q$ are fixed arbitrary Floquet angles;
that is, the prequantization depends on the vector $\chi =(\chi _p,\chi _q)$
whose coordinates are in the range $0\le \chi _q,\chi _q<1$.
It is a well known kinematical restriction \cite{debievre} for torus
quantization that there are
\begin{equation}
\nu ^2N=\nu ^2(2\pi \hbar )^{-1} \label{eq:nh}
\end{equation}
basis states for each degree of freedom, so that $N=(2\pi \hbar )^{-1}$ is
the number of states corresponding to the unit cell .This is a crucial
point: the compactness of the phase space implies in the finiteness of the
dimension of the Hilbert space.
Recalling the translation operators $\widehat{T}_p=\exp \left( \frac{i\alpha
}\hbar \widehat{q}\right) \,$ and $\widehat{T}_q=\exp \left( -\frac{i\beta }%
\hbar \widehat{p}\right) $ that respectively translates momentum by $\alpha $
and position by $\beta $ in the plane, we define minimal translators on the
torus $\widehat{{\bf T}}_{p,\nu ^2N}$ and $\widehat{{\bf T}}_{q,\nu ^2N}$
with their discrete eigenstates $|{\bf q}_n,\nu ^2N>$ and $|{\bf p}_m,\nu
^2N>$ such that
\begin{equation}
\begin{array}{cc}
\widehat{{\bf T}}_{p,\nu ^2N}|{\bf q}_n,\nu ^2N>=e^{\left[ \frac{2\pi i}{\nu
^2N}(n+\chi _q)\right] }|{\bf q}_n,\nu ^2N> & \widehat{{\bf T}}_{p,\nu ^2N}|%
{\bf p}_m,\nu ^2N>=|{\bf p}_{m+1},\nu ^2N> \\
\widehat{{\bf T}}_{q,\nu ^2N}|{\bf q}_n,\nu ^2N>=|{\bf q}_{n+1},\nu ^2N> &
\widehat{{\bf T}}_{q,\nu ^2N}|{\bf p}_m,\nu ^2N>=e^{\left[ -\frac{2\pi i}{%
\nu ^2N}(m+\chi _p)\right] }|{\bf p}_m,\nu ^2N>
\end{array}
\end{equation}
The product of $\nu ^2N$ translations on the basis states $|{\bf q}_n,\nu
^2N>$ or $|{\bf p}_m,\nu ^2N>\,$must return these to the same state, i.e.
these Schwinger operators \cite{schwinger} satisfy
\begin{equation}
\left( \widehat{{\bf T}}_{p,\nu ^2N}\right) ^{\nu ^2N}=e^{2\pi i\chi _q}%
\widehat{{\bf 1}}_{\nu ^2N}^\chi
\end{equation}
and
\begin{equation}
\left( \widehat{{\bf T}}_{q,\nu ^2N}\right) ^{\nu ^2N}=e^{-2\pi i\chi _p}%
\widehat{{\bf 1}}_{\nu ^2N}^\chi \ .
\end{equation}
To define the Hilbert space ${\cal H}_{\nu ^2N}^\chi $, we add the Hermitian
structures
\begin{equation}
<{\bf q}_n,\nu ^2N|{\bf q}_{n^{\prime }},\nu ^2N>=\delta _{n,n^{\prime
}}^{(\nu ^2N)}e^{\frac{2\pi i}{\nu ^2N}(n-n^{\prime })\chi _p},
\label{eq:QQsim}
\end{equation}
and
\begin{equation}
<{\bf p}_m,\nu ^2N|{\bf p}_{m^{\prime }},\nu ^2N>=\delta _{m,m^{\prime
}}^{(\nu ^2N)}e^{-\frac{2\pi i}{\nu ^2N}(m-m^{\prime })\chi _q}.
\label{qbastgp}
\end{equation}
Here we define the {\it $N$-periodic Kronecker delta }
\begin{equation}
\delta _{m,n}^{(N)}\equiv \sum_{j=-\infty }^\infty \delta _{m,n+jN}.
\label{Nkrone}
\end{equation}
The bases are exchanged with the transformation kernel
\begin{equation}
<{\bf p}_m,\nu ^2N|{\bf q}_n,\nu ^2N>=\frac 1{\nu N^{1/2}}e^{2\pi i\frac
1{\nu ^2N}(m+\chi _p)(n+\chi _q)}\equiv F_{m,n}, \label{eq:PQ}
\end{equation}
forming a unitary matrix ( finite Fourier transformation).
Clearly, this last expression allows us to interpret the position ${\bf q}_n$
as corresponding to $q_n=\frac 1{\nu N}(n+\chi _q)$, whereas ${\bf p}_m$
corresponds to $p_m=\frac 1{\nu N}(m+\chi _p)$, leading to
\begin{equation}
<{\bf p}_m,\nu ^2N|{\bf q}_n,\nu ^2N>=(2\pi \hbar )^{-\frac 12}\exp \left(
\frac i\hbar p_mq_n\right) .
\end{equation}
Likewise the $q$-Translator $\widehat{T}_q$, corresponds to a translation in
the plane by $\Delta q=\frac 1{\nu N}$ and the phase change $\exp \left[
2\pi i\chi _q\right] $ results from the translation $\Delta q=\nu $ $\,$%
around the torus. Although the indices $n$ and $m$ can run over all
integers, only $\nu ^2N$ successive values will form a basis for the torus
Hilbert space ${\cal H}_{\nu ^2N}^\chi .$ We may keep to the fundamental
range $\left[ 0,\nu ^2N-1\right] $, corresponding to the square with side $%
\nu $, or extend to the periodic plane, by taking into account the phases $%
\chi $. These considerations apply to each of the $L$ degrees of freedom, so
that in general the fundamental domain is a $\left( 2L\right) $-hypercube.
We see that position and momentum form a discrete web on the torus, as shown
in Fig.~\ref{fig.3.1}, that we call the quantum phase space (QPS), following reference
\cite{galeti1}:
\begin{equation}
x=\left(
\begin{array}{c}
p \\
q
\end{array}
\right) =\frac 1N\left(
\begin{array}{c}
m+\chi _p \\
n+\chi _q
\end{array}
\right) . \label{xtorop}
\end{equation}
\begin{figure}[htb]
\centerline {\epsfxsize=5in \epsffile{fig3-1.eps} }
\caption{ The quantum Phase space for $N=4$. The intersection of the
bold lines in the unit square determine QPS for $\nu =1$, whereas the full
figure corresponds to the choice $\nu =3$.
}
\label{fig.3.1}
\end{figure}
We now consider the relation between the Hilbert spaces of two nested tori $%
{\cal H}_N^\chi $ and ${\cal H}_{\nu ^2N}^{\chi ^{\prime }}$ with $\nu >1.$
The consideration for the QPS corresponding to ${\cal H}_N^\chi $ to be a
sublattice of the QPS of ${\cal H}_{\nu ^2N}^{\chi ^{\prime }}$ is first
that
\begin{equation}
\chi ^{\prime }=\nu \chi -k, \label{nunup}
\end{equation}
where $k=(k_p,k_q)$ is an integer vector that denotes the number of loops
around the torus made by $\chi $ when multiplied by $\nu .$ Thus $\chi
^{\prime }$ is uniquely determined by $\chi .$ The indices for the larger
Hilbert space ${\cal H}_{\nu ^2N}^{\chi ^{\prime }}$ for points in the
fundamental domain of the smaller QPS are given by
\begin{eqnarray}
m^{\prime } &=&\nu m+k_p \\
n^{\prime } &=&\nu n+k_q.
\end{eqnarray}
We can now define the Hilbert space ${\cal H}_N^\chi $ as a projection of
the larger space ${\cal H}_{\nu ^2N}^{\chi ^{\prime }}.$ Indeed, it is easy
to verify that if $|{\bf q}_{n^{\prime }},\nu ^2N>$ are orthogonal position
eigenstates for ${\cal H}_{\nu ^2N}^{\chi ^{\prime }}$, then
\begin{equation}
|{\bf q}_n,N>_\chi =\frac 1{\sqrt{\nu }}\sum_{r=0}^{\nu -1}e^{i2\pi \chi
_pr}|{\bf q}_{\nu n+k}+r,\nu ^2N>_{\chi ^{\prime }} \label{eq:q1q2}
\end{equation}
form an appropriate orthonormal basis for ${\cal H}_N^\chi $. Thus, defining
the projection operator
\begin{equation}
\widehat{{\bf 1}}_N^\chi =\sum_{n=0}^{N-1}|{\bf q}_n,N><{\bf q}_n,N|
\label{uno}
\end{equation}
we verify that this is a Hermitian operator in ${\cal H}_{\nu ^2N}^{\chi
^{\prime }}$, and that
\begin{equation}
\widehat{{\bf 1}}_N^\chi \widehat{{\bf 1}}_N^\chi =\widehat{{\bf 1}}_N^\chi .
\end{equation}
Furthermore the states are obtained as
\begin{equation}
|{\bf \Psi ,}N>=\widehat{{\bf 1}}_N^\chi {\bf |\Psi ,}\nu ^2N>.
\end{equation}
For all operators $\widehat{{\bf A}}_{\nu ^2N}$ acting in ${\cal H}_{\nu
^2N}^{\chi ^{\prime }},$ there is a projected operator which acts on ${\cal H%
}_N^\chi $:
\begin{equation}
\widehat{{\bf A}}_N=\widehat{{\bf 1}}_N^\chi \widehat{{\bf A}}_{\nu ^2N}%
\widehat{{\bf 1}}_N^\chi .
\end{equation}
If $\widehat{{\bf A}}_{\nu ^2N}$ leaves ${\cal H}_N^\chi $ invariant, i.e.
if
\begin{equation}
\lbrack \widehat{{\bf 1}}_N^\chi ,\widehat{{\bf A}}_{\nu ^2N}]=0,
\label{acomu}
\end{equation}
then
\begin{equation}
\widehat{{\bf A}}_N=\widehat{{\bf A}}_{\nu ^2N}\widehat{{\bf 1}}_N^\chi .
\end{equation}
We also verify that, for any pair of operators satisfying (\ref{acomu})
\begin{equation}
\widehat{{\bf A}}_N\widehat{{\bf B}}_N=\widehat{{\bf A}}_{\nu ^2N}\widehat{%
{\bf B}}_{\nu ^2N}\widehat{{\bf 1}}_N^\chi . \label{eq:abpt}
\end{equation}
In particular, we obtain the Schwinger translation operators for ${\cal H}%
_N^\chi $ as the projection of those in ${\cal H}_{\nu ^2N}^{\chi ^{\prime
}} $:
\begin{equation}
\widehat{{\bf T}}_{q,N}=\widehat{{\bf T}}_{q,\nu ^2N}\widehat{{\bf 1}}%
_N^\chi \qquad ;\qquad \widehat{{\bf T}}_{p,N}=\widehat{{\bf T}}_{p,\nu ^2N}%
\widehat{{\bf 1}}_N^\chi .
\end{equation}
Let us now take the limit $\nu \rightarrow \infty $. Clearly, the variable $%
q_n$ becomes continuous in this limit as the volume of the torus $\nu
^{2L}\rightarrow \infty $.Throughout the limit, the relation (\ref{nunup})
defines an appropriate $\chi ^{\prime }\left( \chi \right) $. The main step
to recover the Banach space ${\cal H}_{{\Bbb R}}$ for the plane $\left(
2L\right) $-dimensional phase space is to redefine the normalization so that
\begin{equation}
<q|q^{\prime }>=\delta (q-q^{\prime }), \label{qbastg2}
\end{equation}
introducing the Dirac delta function on the right, and the continuous
Fourier integral for the change of basis:
\begin{equation}
<p|\psi >=\left( 2\pi \hbar \right) ^{-L/2}\int dq\exp \left( \frac{ipq}%
\hbar \right) <q|\psi >
\end{equation}
In all other respects, the kinematics in the plane will coincide with that
of an infinitely large torus. However, the normalization condition (\ref
{qbastg2}) implies a change in the way we express the sates in ${\cal H}%
_N^\chi $ in terms of those of ${\cal H}_{{\Bbb R}}$. For that purpose, we
recall that for an unit torus
\begin{equation}
\sum_{r=0}^{N-1}e^{i\frac{2\pi }N(m-n)r}=N\delta _{m,n}^{(N)},
\label{eq:114}
\end{equation}
so we extend the definition of the $N$-periodic Kroeneker delta function to
real numbers $x$ and $y$:
\begin{equation}
\delta _{x,y}^{(N)}\equiv \left\langle e^{i\frac{2\pi }N\left( x-y\right)
k}\right\rangle _k, \label{eq:kroneinf}
\end{equation}
where $\left\langle ...\right\rangle _k$ denotes the average over $k,$
\begin{equation}
\left\langle ...\right\rangle _k=\lim_{r\rightarrow \infty }\frac
1r\sum_{k=-\frac r2}^{\frac r2}. \label{kprom}
\end{equation}
From (\ref{eq:kroneinf}) we see that $\delta _{x,y}^{(N)}$ only depends on
the difference $x-y$, so let us take $y=0$ for simplicity. For $x=r$ an
integer number, the argument in (\ref{eq:kroneinf})\ has period $N$. Then,
the average is just (\ref{eq:114}) divided by $N,$ so the definition (\ref
{eq:kroneinf})\ is consistent with (\ref{Nkrone}) for integer arguments. Let
us now suppose $x=\frac rd,$ a rational number. The argument in (\ref
{eq:kroneinf})\ thus has period $Nd$, so that
\begin{equation}
\delta _{x,0}^{(N)}=\frac 1{Nd}\sum_{k=0}^{Nd-1}e^{i\frac{2\pi }N\frac
rdk}=\frac 1{Nd}\frac{1-e^{i2\pi r}}{1-e^{i\frac{2\pi }{Nd}r}}=\left\{
\begin{array}{cc}
1 & \mbox{ if }r=0\qquad \mbox{mod}(Nd) \\
0 & \mbox{ otherwise}
\end{array}
\right. . \label{krorac}
\end{equation}
Hence, once more the $N$-periodic Kroeneker delta function is different from
zero only for $x$ being zero modulo $N$. By allowing $d\rightarrow \infty $
in (\ref{krorac}) we can extend the definition of $\delta _{x,y}^{(N)}$ to
irrational numbers, so that
\begin{equation}
\delta _{x,y}^{(N)}=\left\{
\begin{array}{cc}
1 & \mbox{ only if }\left( x-y\right) =0\qquad \mbox{mod}(N) \\
0 & \mbox{ otherwise}
\end{array}
\right. .
\end{equation}
The definition (\ref{eq:kroneinf}) is an interpolation of (\ref{Nkrone}),
which will allows us to perform, not only sums with the $N$ periodic
Kroeneker delta function, but also integrals. Indeed, we will relie on the
formal equivalence:
\begin{equation}
\delta _{x,y}^{(N)}=\left\langle \delta (\frac{x-y}N-k)\right\rangle _k.
\label{eq:dirkro}
\end{equation}
This is a consequence of the definition(\ref{eq:kroneinf}) and the Poisson
sum formula,
\begin{equation}
\sum_{t\in {\Bbb Z}}\delta (x-t)=\sum_{k\in {\Bbb Z}}e^{i2\pi kx}.
\label{eq:poisson}
\end{equation}
Indeed, from the definition (\ref{eq:kroneinf}) we have
\begin{equation}
\frac 1\nu \sum_{k=-\frac \nu 2}^{\frac \nu 2}e^{i\frac{2\pi }N\left(
x-y\right) k}=\frac 1\nu \sum_{k=-\frac \nu 2}^{\frac \nu 2}\delta (\frac{x-y%
}N-k)+\frac 1\nu R_\nu (x-y)
\end{equation}
with
\begin{equation}
R_\nu (x-y)=\sum_{k=-\frac \nu 2}^{\frac \nu 2}e^{i\frac{2\pi }N\left(
x-y\right) k}-\sum_{k=-\frac \nu 2}^{\frac \nu 2}\delta (\frac{x-y}N-k).
\end{equation}
From the Poisson sum formula (\ref{eq:poisson}), we have that, $\lim_{\nu
\rightarrow \infty }R_\nu (x-y)=0,$ upon an appropriate ordering of the
limits $\nu \rightarrow \infty $ and the width of the delta function$%
\rightarrow 0$.A\ consequence of (\ref{eq:dirkro}) is that, for any function
$f(t),$
\begin{equation}
\left\langle \delta (\frac{x-y}N-t)f(t)\right\rangle _t=\delta
_{x,y}^{(N)}\;f(\frac{x-y}N). \label{eq:dirkrox}
\end{equation}
Changing the origin, so as to keep the unit torus at the center of the
larger torus, (\ref{eq:q1q2}) must be replaced by
\begin{equation}
|{\bf q}_n,N>=\lim_{\nu \rightarrow \infty }\frac 1\nu \sum_{r=-\frac \nu
2}^{\frac \nu 2}|\frac{n+\chi _q}N+r>e^{2\pi ir\chi _p}. \label{qpltor}
\end{equation}
A straightforward calculation using (\ref{eq:dirkrox}) shows that the
orthonormality conditions (\ref{eq:QQsim}) are obtained for the states
defined in (\ref{qpltor}) with the normalization (\ref{qbastg2}). So we
consider that the states and the operators in the unit torus are obtained
from the plane by projections. Recalling the definition (\ref{kprom}) of the
average, (\ref{qpltor}) can be written as
\begin{equation}
|{\bf q}_n,N>=\left\langle |\frac{n+\chi _q}N+r>e^{2\pi ir\chi
_p}\right\rangle _r \label{eq:qQ}
\end{equation}
and the projection operator $\widehat{{\bf 1}}_N^\chi \,$ is then given by (%
\ref{uno}). In the case of $\chi =0\,$ we thus retrieve the definition of
Hannay and Berry \cite{hanay} for ${\cal H}_N^0$ as the average over a
periodic array of Dirac delta distributions. We will now derive the Weyl
representation on the torus by projecting the properties that have been well
established in the plane.
\section{The Weyl representation on the plane}
\setcounter{equation}{0}
We here summarize the results obtained for the plane in reference \cite
{ozrep} that will be projected onto the torus in the following sections. We
define the operator corresponding to a general translation in phase space by
the $\left( 2L\right) $-dimensional vector $\xi =(\xi _p,\xi _q)$ as
\begin{equation}
\hat{T}_\xi \equiv \exp \left( \frac i\hbar \xi \wedge \hat{x}\right) \equiv
\exp \left[ \frac i\hbar (\xi _p.\hat{q}-\xi _q.\hat{p})\right] ,
\label{eq:tcor}
\end{equation}
where naturally $\hat{x}=(\hat{p},\hat{q})$ and the symplectic product $\xi
\wedge \eta $ is defined as
\begin{equation}
\xi \wedge \eta =({\frak J}\xi ).\eta
\end{equation}
with
\begin{equation}
{\frak J}=\left[
\begin{array}{c|c}
0 & -1 \\ \hline
1 & 0
\end{array}
\right] .
\end{equation}
$\hat{T}_\xi $ is also known as a {\it Heisenberg operator}. In the case
where either $\xi _p$ or $\xi _q=0$, we obtain respectively the operators $%
\hat{T}_q$ or $\hat{T}_p$ mentioned in the preceding section.
Acting on the Banach space ${\cal {H}_{{\Bbb R}}}$ we have
\begin{equation}
\widehat{T}_\xi |q_a>=e^{\frac i\hbar \xi _p(q_a+\frac{\xi _q}2)}|q_a+\xi _q>
\label{eq:tq}
\end{equation}
and
\begin{equation}
\widehat{T}_\xi |p_a>=e^{-\frac i\hbar \xi _q(p_a+\frac{\xi _p}2)}|p_a+\xi
_p>.
\end{equation}
The classical group property is maintained within a phase factor:
\begin{equation}
\hat{T}_{\xi _2}\hat{T}_{\xi _1}=\hat{T}_{\xi _1+\xi _2}\ \exp [\frac{-i}{%
2\hbar }\xi _1\wedge \xi _2]=\hat{T}_{\xi _1+\xi _2}\ \exp [\frac{-i}\hbar
D_3(\xi _1,\xi _2)], \label{eq:tt}
\end{equation}
where $D_3$ is the symplectic area of the triangle determined by two of its
sides. Evidently, the inverse of the unitary operator $\hat{T}_\xi ^{-1}=%
\hat{T}_\xi ^{\dag }=\hat{T}_{-\xi }$ and we can generalize (\ref{eq:tt}):
\begin{equation}
\widehat{T}_{\xi _1}...\widehat{T}_{\xi j}=\widehat{T}_{\xi _1+....+\xi
_j}e^{\frac i\hbar D_{j+1}(\xi _1,....\xi _j)}, \label{eq:ttt}
\end{equation}
where $D_{j+1}(\xi _1,....\xi _j)$ denotes the symplectic area of the $%
\left( j+1\right) $ sided polygon formed by the chords, $(\xi _1,....\xi _j)$%
.
The operator corresponding to a phase space reflection about a point $%
x=(p,q) $ is \cite{ozrep}
\begin{equation}
\widehat{R}_x\equiv (4\pi \hbar )^{-L}\int d\xi \quad e^{\frac i\hbar
x\wedge \xi }\widehat{T}_\xi . \label{eq:rint}
\end{equation}
This operator has the following properties \cite{ozrep} :
\begin{equation}
\widehat{T}_\xi =(\pi \hbar )^{-L}\int dx\quad e^{-\frac i\hbar x\wedge \xi }%
\widehat{R}_x\ , \label{eq:tintr}
\end{equation}
\begin{equation}
\widehat{R}_x\widehat{T}_\xi =\widehat{R}_{x-\xi /2}e^{-\frac i\hbar x\wedge
\xi }\ , \label{eq:rt}
\end{equation}
\begin{equation}
\widehat{T}_\xi \widehat{R}_x=\widehat{R}_{x+\xi /2}e^{-\frac i\hbar x\wedge
\xi }\ \label{eq:tr}
\end{equation}
and
\begin{equation}
\widehat{R}_{x_1}\widehat{R}_{x_2}=\widehat{T}_{2(x_2-x_1)}e^{\frac i\hbar
2x_1\wedge x_2}, \label{eq:rr}
\end{equation}
so that
\begin{equation}
\widehat{R}_x\widehat{R}_x=\widehat{1}\ .
\end{equation}
The trace of the translation is
\begin{eqnarray}
Tr\hat{T}_\xi &=&\int <q|T_\xi |q>dq=(2\pi \hbar )^L\delta (\xi _p)\delta
(\xi _q) \nonumber \\
&=&(2\pi \hbar )^L\delta (\xi )\
\end{eqnarray}
and then taking the Fourier transform,
\begin{equation}
Tr\hat{I\!\!\!\!R}_x=Tr2^L\hat{R}_x=(2\pi \hbar )^{-L}\int d\xi \exp \left[
\frac i\hbar \ x\wedge \xi \right] Tr\hat{T}_\xi =1\ ,
\end{equation}
where it is now also convenient to define the exact Fourier transform $\hat{%
I\!\!\!\!R}_x$ of $\hat{T}_\xi $. We recall that the classical
transformation $R_x$ has a single fixed point ($x$ itself), whereas $T_\xi $
has fixed points only if $\xi =0$, when all points are fixed. These results
are in general agreement with our intuition as to the classical
correspondence of the traces of unitary operators. i.e. that the trace is
related to the classical fixed points.
The above properties allow any operator $\hat{A}$ to be expressed as a
linear superposition of elementary translation operators:
\begin{equation}
\hat{A}=\int \frac{d\xi }{(2\pi \hbar )^L}\ A(\xi )\hat{T}_\xi \ .
\label{eq:cora}
\end{equation}
The confirmation results from
\begin{equation}
Tr(\hat{T}_{-\xi }\hat{A})=A(\xi )\ , \label{eq:acor}
\end{equation}
The analogy with the classical chord generating function for canonical
transformations is discussed in \cite{ozrep}. We can equally represent any
operator $\hat{A}$ as a superposition of reflections:
\begin{equation}
\hat{A}=\int \frac{dx}{(2\pi \hbar )^L}\ A(x)\hat{I\!\!\!\!R}_x=\int \frac{dx%
}{(\pi \hbar )^L}\ A(x)\hat{R}_x\ . \label{eq:cena}
\end{equation}
Again we obtain the expansion coefficient by calculating
\begin{equation}
Tr(\hat{I\!\!\!\!R}_x\hat{A})=A(x)\ . \label{eq:acen}
\end{equation}
Notice that comparison of (\ref{eq:cora}) and (\ref{eq:cena}) with (\ref
{eq:rint}) and (\ref{eq:tintr}) yields
\begin{equation}
R_x(\xi )=2^{-L}\exp \left[ \frac i\hbar \ x\wedge \xi \right] \ \ \ %
\mbox{and}\ \ \ T_\xi (x)=\exp \left[ -\frac i\hbar \ x\wedge \xi \right] \ .
\end{equation}
In analogy with our previous result, we may refer to $A(x)$ as the {\it %
center representation} of the operator $\hat{A}$, but the historic term is
the {\it Weyl representation}.
The product laws of center and chord representations are of fundamental
interest. Starting with the chord representation, we have, for the product $%
\hat{A}_n\hat{A}_{n-1}\cdots \hat{A}_1$,
\begin{eqnarray}
&&A_n.A_{n-1}\cdots A_1(\xi )=\left( \frac 1{2\pi \hbar }\right)
^{L(n-1)}\int d\xi _n\cdots d\xi _1A_n(\xi _n)\cdots \nonumber \\
&&A_1(\xi _1)\delta (\xi _1+\cdots \xi _n-\xi )\exp \left[ -\frac i\hbar
D_{n+1}(\xi _1,\cdots ,\xi _n)\right] , \label{eq:ancor}
\end{eqnarray}
where we note that the Dirac $\delta $-function has reduced the $(n+2)$%
-sided polygon with symplectic area $D_{n+2}$ to an $(n+1)$-sided polygon,
with $n$ free sides. Evidently, we can now use the $\delta $-function to
remove one of the variables in the integral, but (\ref{eq:ancor}) is in its
most symmetric form.
We shall also need integral formulae for the product of operators in the
center representation. The result depends crucially on the parity of the
number of operators\cite{ozrep}, so we will start with the simplest case
where $n=2$. Proceeding from the definition ~(\ref{eq:cena}), we obtain
\begin{equation}
A_2.A_1(x)=\left( \frac 1{\pi \hbar }\right) ^{2L}\int
dx_2dx_1A_2(x_2)A_1(x_1)\exp \left[ \frac i\hbar \Delta _3(x,x_1,x_2)\right]
. \label{eq:aacen}
\end{equation}
where $\Delta _3(x,x_1,x_2)=2(x_1\wedge x_2+x_2\wedge x+x\wedge x_1)$ is the
symplectic area of the triangle whose midpoints are, $x,x_1,x_2$.
The extension to $(2n)$ operators is \cite{ozrep}
\begin{eqnarray}
&&A_{2n}\cdots A_1(x)= \nonumber \\
&&\left( \frac 1{\pi \hbar }\right) ^{2nL}\int dx_{2n}\cdots
dx_1A_{2n}(x_{2n})\cdots A_1(x_1)\exp \left\{ \frac i\hbar \Delta
_{2n+1}(x,x_1,\cdots ,x_{2n})\right\} . \label{eq:ancen}
\end{eqnarray}
Here the symplectic area $\Delta _{2n+1}$ corresponds to the $(2n+1)$-sided
polygon circumscribed around the centers $(x,x_1,\cdots ,x_{2n})$.
The main advantage of the chord and center representation is their
symplectic invariance. It is well known that linear classical canonical
transformations $x^{\prime }=Mx$ correspond to unitary transformations in $%
{\cal {H}_{{\Bbb R}}}$%
\begin{equation}
\widehat{A}\rightarrow \widehat{U}_M\widehat{A}\widehat{U}_M^{-1}.
\end{equation}
The effect of such a unitary transformation on the chord and center
representation is merely
\begin{equation}
A(x)\rightarrow A(Mx)\quad \mbox{and}\quad A(\xi )\rightarrow A(M\xi ).
\end{equation}
\section{Weyl representation in the torus}
\setcounter{equation}{0}
\subsection{ Translation and Reflection Operators on the torus}
In this section we will project the translations $\widehat{T}_\xi $ and
reflections $\widehat{R}_x$ operators defined on ${\cal H}_{{\Bbb R}}$ onto
the Hilbert space ${\cal H}_N^\chi $. Again we will treat the case of one
degree of freedom explicitly, since the generalization is obvious. For this
purpose we first investigate the action of the translation operators $%
\widehat{T}_\xi $ on the $|{\bf q}_n,N>$ basis vectors defined by (\ref
{eq:qQ}). From the effect of a translation (\ref{eq:tq}) on a single
position in the plane, we have
\begin{equation}
\widehat{T}_\xi |{\bf q}_n,N>=\left\langle |\frac{n+\chi _q}N+k+\xi
_q>e^{2\pi ik\chi _p}e^{\frac i\hbar \xi _p(\frac{n+\chi _q}N+k+\frac 12\xi
_q)}\right\rangle _k,
\end{equation}
using the relation (\ref{eq:nh}) between $N$ and $\hbar $. This vector will
belong to ${\cal H}_N^\chi $ only if it has the form (\ref{eq:qQ}), i.e., if
we can write $\xi _q=\frac sN$ with $s$ an integer.
A similar treatment in the $|{\bf p}_m,N>$ representation implies that
\begin{equation}
\widehat{T}_\xi |{\bf p}_m,N>=\left\langle |\frac{m+\chi _p}N+k+\xi
_p>e^{-2\pi ik\chi _q}e^{-\frac i\hbar \xi _q(\frac{m+\chi _q}N+k+\frac{\xi
_p}2)}\right\rangle _k,
\end{equation}
which does not belong to ${\cal H}_N^\chi $ unless we can write $\xi
_p=\frac rN$ with $r$ an integer. So, as was already pointed in \cite
{debievre}, the only translations that leave ${\cal H}_N^\chi $ invariant
are those whose chords are $\xi =(\frac rN,\frac sN)$. For these cases we
have
\begin{eqnarray}
\widehat{T}_\xi |{\bf q}_n,N &>&=\left\langle |\frac{n+\chi _q}N+k+\frac
sN>e^{2\pi ik\chi _p}e^{i2\pi N\frac rN(\frac{n+\chi _q}N+k+\frac
s{2N})}\right\rangle _k \nonumber \\
&=&e^{\frac{i2\pi }Nr(n+\frac s2+\chi _q)}\left\langle |\frac{n+s+\chi _q}%
N+k>e^{2\pi ik(\chi _p+r)}\right\rangle _k \nonumber \\
&=&e^{\frac{i2\pi }Nr(n+\frac s2+\chi _q)}|{\bf q}_{n+s},N>. \label{eq:tQ}
\end{eqnarray}
In short, we obtain the torus operator $\widehat{{\bf T}}_\xi ^\chi $ in
terms of the plane operator $\widehat{T}_\xi $ as
\begin{eqnarray}
\widehat{{\bf 1}}_N^\chi \widehat{T}_\xi \widehat{{\bf 1}}_N^\chi &=&\left\{
\begin{tabular}{lll}
$\widehat{{\bf T}}_\xi ^\chi $ & $\mbox{if }$ $\xi =\left( \frac rN,\frac
sN\right) ,$where $r$ and $s$ are integers & \\
$0$ & otherwise &
\end{tabular}
\right. \\
&=&\widehat{{\bf T}}_\xi ^\chi \left\langle \delta \left( \xi -\frac
1N\left( r,s\right) \right) \right\rangle _{\left( r,s\right) }=\widehat{T}%
_\xi \widehat{{\bf 1}}_N^\chi \ , \label{projt}
\end{eqnarray}
where the torus translation operators $\widehat{{\bf T}}_\xi ^\chi \equiv
\widehat{{\bf T}}_{r,s}^\chi $ are defined through
\begin{equation}
\widehat{{\bf T}}_{r,s}^\chi |{\bf q}_n,N>=e^{i\frac{2\pi }Nr(n+\chi
_q+s/2)}|{\bf q}_{n+s},N> \label{eq:TQ}
\end{equation}
and
\begin{equation}
\widehat{{\bf T}}_{r,s}^\chi |{\bf p}_m,N>=e^{-i\frac{2\pi }Ns(m+\chi
_p+r/2)}|{\bf p}_{m+r},N>. \label{eq:TP}
\end{equation}
The last equality in (\ref{projt}) holds because $\widehat{T}_\xi $ and $%
\widehat{{\bf 1}}_N^\chi $ commute for $\xi =\left( \frac rN,\frac sN\right)
$. We then see that the only translation operators that do not vanish on
projection to the torus are those that leave ${\cal H}_N^\chi $ invariant.
They correspond precisely to those classical transformations that preserve
the QPS web. To simplify the notation, we will usually assume implicitly the
$\chi $ dependence.
Let us now study some properties of the torus translation operators. For the
case where the chords are the minimal translations in any one of the $q$ or $%
p$ directions, we recover the Schwinger operators \cite{schwinger} so that,
\begin{equation}
\widehat{{\bf T}}_{0,1}|{\bf q}_n,N>\equiv \widehat{{\bf T}}_q|{\bf q}_n,N>=|%
{\bf q}_{n+1},N>
\end{equation}
and,
\begin{equation}
\widehat{{\bf T}}_{1,0}|{\bf p}_m,N>\equiv \widehat{{\bf T}}_p|{\bf p}_m,N>=|%
{\bf p}_{m+1},N>.
\end{equation}
The kernel (\ref{eq:PQ}) implies that,
\begin{equation}
\widehat{{\bf T}}_p\widehat{{\bf T}}_q=e^{-\frac{2\pi i}N}\widehat{{\bf T}}_q%
\widehat{{\bf T}}_p. \label{eq:grupo}
\end{equation}
so any translation operator in ${\cal H}_N^\chi $ is defined as
\begin{equation}
\widehat{{\bf T}}_\xi \equiv \widehat{{\bf T}}_{r,s}=e^{-\frac{i\pi rs}N}%
\widehat{{\bf T}}_p^r\widehat{{\bf T}}_q^s\ , \label{eq:Tcor}
\end{equation}
with chords $\xi =(\frac rN,\frac sN).$ We can express the matrix elements
of the translation operators in the $|{\bf q}_n,N>$ basis,
\begin{equation}
<{\bf q}_m,N|\widehat{{\bf T}}_{r,s}|{\bf q}_n,N>=e^{-i\frac{2\pi }Nr(\frac{%
m+n}2+\chi _q)}\delta _{m,n+s}^{(N)}e^{i\frac{2\pi }N(\frac r2+\chi
_p)(m-n-s)}, \label{eq:TQQ}
\end{equation}
using the orthonormality relations of the states (\ref{eq:QQsim}).The fact
that the Hilbert space has finite dimension implies that linear operators
acting on it will be represented by $N\times N$ matrices. Then $N^2$
linearly independent matrices will form a basis for the operators in ${\cal H%
}_N^\chi $. This is clear from the symmetries of the translation operators;
through their action on ${\cal H}_N^\chi $ (\ref{eq:TQ}) we see that
\begin{equation}
\widehat{{\bf T}}_{\xi +{\bf k}}=(-1)^{sk_p-rk_q+k_pk_qN}e^{i2\pi (k_p\chi
_q-k_q\chi _p)}\widehat{{\bf T}}_\xi =e^{-i2\pi N\left[ \left( \frac \xi
2+\frac \chi N\right) \wedge {\bf k}+\frac 14{\bf k}\widetilde{{\frak J}}%
{\bf k}\right] }\widehat{{\bf T}}_\xi , \label{eq:TTsim}
\end{equation}
where ${\bf k}=(k_p,k_q)$ is a vector with integer components denoting
chords that perform respectively $k_p$ and $k_q$ loops around the
irreducible circuits of the torus. We have also defined the symmetric matrix
\begin{equation}
\widetilde{{\frak J}}=\left[
\begin{array}{c|c}
0 & 1 \\ \hline
1 & 0
\end{array}
\right] .
\end{equation}
If we perform ${\bf k}$ loops around the torus, (\ref{eq:TTsim}) implies
that we recover the identity operator only up to a phase:
\begin{equation}
\widehat{{\bf T}}_{{\bf k}}=(-1)^{k_pk_qN}e^{i2\pi (k_p\chi _q-k_q\chi _p)}%
\widehat{{\bf 1}}_N^\chi . \label{tNQ}
\end{equation}
Thus, to have a basis of operators we only need $r$ and $s$ in the range $%
[0,N-1]$, that is, we only need translations that perform less than one loop
around the torus.
The second phase factor in expression (\ref{eq:TTsim}) comes from the Bloch
boundary conditions, but the $(-1)^{sk_p-rk_q+k_pk_qN}$ factor shows that we
need two loops around the torus to recover the same operator, doubling the
expected periodicity. This will have crucial importance in the construction
of the reflection operators.
An important property of the translation operators, which can be deduced
from (\ref{eq:grupo}) is
\begin{equation}
\widehat{{\bf T}}_{\xi _2}\widehat{{\bf T}}_{\xi _1}=\widehat{{\bf T}}_{\xi
_1+\xi _2}e^{i\pi N\xi _1\wedge \xi _2}, \label{eq:TT}
\end{equation}
which generalizes to
\begin{equation}
\widehat{{\bf T}}_{\xi _1}...\widehat{{\bf T}}_{\xi j}=\widehat{{\bf T}}%
_{\xi _1+....+\xi _j}e^{i\pi ND_{j+1}(\xi _1,....\xi _j)},
\end{equation}
where $D_{j+1}(\xi _1,....\xi _j)$ denotes the area of $j+1$ sided polygon
formed by the chords, $(\xi _1,....\xi _j)$, exactly as (\ref{eq:ttt}) in
the plane case. From (\ref{eq:TT}) and the unitarity of $\widehat{{\bf T}}%
_\xi ,$ we can see that
\begin{equation}
\widehat{{\bf T}}_\xi ^{\dag }=\widehat{{\bf T}}_\xi ^{-1}=\widehat{{\bf T}}%
_{-\xi }. \label{eq:T-1}
\end{equation}
Notice that (\ref{eq:TT}) reduces to (\ref{eq:grupo}) for the particular case that $\xi _1$ and $\xi _2$ are vectors along the coordinate axis.
For the reflection operators, we use their definition (\ref{eq:rint}) in
terms of plane translations and then (\ref{projt}) projects the translations
onto the torus, so that
\begin{eqnarray}
\widehat{{\bf R}}_x^\chi &=&\widehat{{\bf 1}}_N^\chi \widehat{R}_x\widehat{%
{\bf 1}}_N^\chi =(4\pi \hbar )^{-L}\int d\xi \quad e^{\frac i\hbar x\wedge
\xi }\widehat{{\bf T}}_\xi \left\langle \delta \left( \xi -\frac{{\bf j}}%
N\right) \right\rangle _{{\bf j}} \nonumber \\
&=&(4\pi \hbar )^{-L}\left\langle e^{i2\pi x\wedge {\bf j}}\widehat{{\bf T}}%
_{\frac{{\bf j}}N}\right\rangle _{{\bf j}}.
\end{eqnarray}
To perform the average we use the periodicity of the torus translation (\ref
{eq:TTsim}), but we perform two loops around the torus, so that
\begin{eqnarray}
\widehat{{\bf R}}_x^\chi &=&(4\pi \hbar )^{-L}\left( \frac 1{2N}\right)
^2\sum_{r=0}^{2N-1}\sum_{s=0}^{2N-1}\left\langle e^{-i2\pi N\left[ \left(
\frac{{\bf j}}{2N}+\frac \chi N\right) \wedge 2{\bf k}+{\bf k}\widetilde{%
{\frak J}}{\bf k}\right] }e^{i2\pi x\wedge ({\bf j+}2{\bf k}N{\bf )}}%
\widehat{{\bf T}}_{\frac{{\bf j}}N}\right\rangle _{{\bf k}} \nonumber \\
&=&\frac 1{2N}\sum_{r=0}^{2N-1}\sum_{s=0}^{2N-1}e^{-i2\pi x\wedge {\bf j}}%
\widehat{{\bf T}}_{\frac{{\bf j}}N}\left\langle e^{-i2\pi \left[ \left( \chi
-xN\right) \wedge 2{\bf k}\right] }\right\rangle _{{\bf k}}.
\end{eqnarray}
The ${\bf k}$ average is different from zero only if the point $x$ is such
that $x=x_{a,b}\equiv \frac 1N\left(
\begin{array}{c}
a+\chi _p \\
b+\chi _q
\end{array}
\right) ,\,$ with $a\,$ and $b$ half integers.
In short,
\begin{eqnarray}
\widehat{{\bf 1}}_N^\chi \widehat{R}_x\widehat{{\bf 1}}_N^\chi &=&\left\{
\begin{tabular}{lll}
$\widehat{{\bf R}}_{x_{a,b}}^\chi $ & $\mbox{if }$ $x=\left( \frac{a+\chi _p}%
N,\frac{b+\chi _q}N\right) ,$ where $a$ and $b$ are half-integers & \\
$0$ & otherwise &
\end{tabular}
\right. \\
&=&\widehat{{\bf R}}_{x_{a,b}}^\chi \left\langle \delta \left( x-\frac{{\bf k%
}}{2N}+\frac \chi N\right) \right\rangle _{{\bf k}}=\widehat{R}_x\widehat{%
{\bf 1}}_N^\chi , \label{rtoro}
\end{eqnarray}
where
\begin{equation}
\widehat{{\bf R}}_{x_{a,b}}^\chi =\frac
1{2N}\sum_{r=0}^{2N-1}\sum_{s=0}^{2N-1}e^{i2\pi Nx\wedge \xi }\widehat{{\bf T%
}}_\xi ^\chi \label{eq:RRT}
\end{equation}
is the torus reflection on the center point $x_{a,b}$. The last equality in (%
\ref{rtoro}) holds because $\widehat{{\bf R}}_{x_{a,b}}^\chi $ commutes with
$\widehat{{\bf 1}}_N^\chi .$ Again, we will usually omit the explicit $\chi
$ dependence.
The construction of the reflection operators on the torus replaces the
Fourier transform (\ref{eq:rint}) by a Fourier sum on the torus translation
operators. The sums are to be taken over operators on one complete period;
the symmetry properties (\ref{eq:TTsim}) show that this period is obtained
with chords that perform two loops around the torus, i.e. , the period is
double that expected. So, although the basis of operators is formed with
chords that perform up to one loop around the torus in the Fourier sum, we
have to sum over chords that perform up to two loops. Thus, the basis
operators are summed twice, but with different Fourier phases.
In what follows, the subscripts $(a,b)$ for the discrete centers and $(r,s)$
for the lattice of chords are implicit and they will be explicitly written
only to avoid possible confusion. With the use of (\ref{eq:114}) we can
derive the following extensively used relations,
\begin{equation}
\sum_{a=0}^{N-1/2}\sum_{b=0}^{N-1/2}e^{i2\pi Nx\wedge \xi }=(2N)^2\delta
_{r,0}^{(2N)}\delta _{s,0}^{(2N)}, \label{eq5}
\end{equation}
where all the sums over $a$ and $b$ are taken with step $\frac 12$, and
\begin{equation}
\sum_{r=0}^{2N-1}\sum_{s=0}^{2N-1}e^{i2\pi Nx\wedge \xi }=(2N)^2\delta
_{b,0}^{(N)}\delta _{a,0}^{(N)}. \label{eq6}
\end{equation}
Here $\delta _{b,0}^{(2N)}$ is a period-$2N$ Kronecker delta function.
Inserting (\ref{eq:TQ}) in (\ref{eq:RRT})we find the action of the
reflection operators on the Hilbert space:
\begin{equation}
\widehat{{\bf R}}_x|{\bf q}_n,N>=e^{i\frac{2\pi }N2(b-n)(a+\chi _p)}|{\bf q}%
_{2b-n},N> \label{eq:RQ}
\end{equation}
and
\begin{equation}
\widehat{{\bf R}}_x|{\bf p}_m,N>=e^{i\frac{2\pi }N2(a-m)(b+\chi _q)}|{\bf p}%
_{2a-m},N>. \label{eq:RP}
\end{equation}
The unitarity of $\widehat{{\bf R}}_x$ is ensured by (\ref{eq:RQ}) and (\ref
{eq:RP}).
We then see that $\widehat{{\bf R}}_x$ reflects the (QPS) web about the
point $x=(\frac{a+\chi _p}N,\frac{b+\chi _q}N)$. We need to include
half-integer values of $a$ and $b$ so that with a given $|{\bf q}_n,N>$ we
can span all ${\cal H}_N^\chi $ by applying different $\widehat{{\bf R}}_x$.
This is in complete agreement with the fact that the reflections that leave
invariant the web formed by the QPS must include half-integer values of $a$
and $b$, conferring on these half-integers a clear geometrical meaning. So
the centers of the reflections form a web whose spacing is half that of the
QPS, as shown in Fig.~\ref{fig.3.2}. Once more, the only operators that do not vanish on
projection to the torus are those that leave ${\cal H}_N^\chi $ invariant.
These correspond classically to those transformations that leave the QPS web
invariant.
\begin{figure}[htb]
\centerline {\epsfxsize=3in \epsffile{fig3-2.eps} }
\caption{ The quantum Phase space QPS for $N=4$ (solid line). The Weyl
phase space WPS (doted line) is the grid of points $x$, centers of the
reflection operators in QPS. The area in the bold square is the
quarter-torus in which lie the centers $x,$ which label the basis of
operators ${\bf R}_x$.
}
\label{fig.3.2}
\end{figure}
The matrix elements of the reflection operators in the $|{\bf q}_n,N>$ basis
are
\begin{equation}
<{\bf q}_m,N|\widehat{{\bf R}}_x|{\bf q}_n,N>=e^{i\frac{2\pi }N(m-n)(a+\chi
_q)}\delta _{m,2b-n}^{(N)}e^{i\frac{2\pi }Na(2b-n-m)}. \label{eq:RQQ}
\end{equation}
From (\ref{eq:RQ}) we can see the symmetry properties of these operators,
\begin{equation}
\widehat{{\bf R}}_{x+\frac{{\bf k}}2}=(-1)^{bk_p+ak_q+k_pk_qN}\widehat{{\bf R%
}}_x \label{eq:symR}
\end{equation}
where ${\bf k}=(k_p,k_q)$ is a vector with integer components. It is
important to see that the domain of the variables $a$ and $b$ being integer
and half-integer values, we have $(2N)^2$ different
reflection operators in the unit square. But the symmetry properties (\ref{eq:symR}) show that
only $N^2$ of them are independent, so we take the values of $a$ and $b$
that belong to $[0,\frac{N-1}2]$ ; this forms a complete set of independent
operators. That is, only one quarter of the torus is needed to define a
complete set of reflection operators. The values of $x=(\frac{a+\chi _p}N,%
\frac{b+\chi _q}N)$ generated by these values of $a$ and $b$ do not all
belong to the QPS; indeed we define here another space, the {\it Weyl phase
space}, WPS, formed by the support of $x$ this is shown by the bold face
area in Fig.~\ref{fig.3.2}. In the case where $N$ is odd, we will see later that WPS can
be defined such that it coincides with the QPS.
By the use of (\ref{eq5}) we can see that,
\begin{equation}
\widehat{{\bf T}}_\xi =\frac
1{2N}\sum_{a=0}^{N-1/2}\sum_{b=0}^{N-1/2}e^{-i2\pi Nx\wedge \xi }\widehat{%
{\bf R}}_x, \label{eq:tsumr}
\end{equation}
where we are again taking the sum with the indices running in an interval
twice as large as that needed to define a basis of operators. This is so
because (\ref{eq:symR}) implies that classically equivalent reflections,
through points diametrically opposed on any of the circuits of the torus,
are only equal up to a phase.
We now investigate the group or cocycle properties of the translations and
reflections defined in this section. It is important to note that the
transformations treated here are such that they leave the web formed by the
QPS invariant at the classical level, as well as the Hilbert space ${\cal H}%
_N^\chi $. With the help of (\ref{eq:tsumr}) and (\ref{eq:TT}), we obtain
the following properties for these operators,
\begin{equation}
\widehat{{\bf R}}_x\widehat{{\bf T}}_\xi =\widehat{{\bf R}}_{x-\xi
/2}e^{-i2\pi Nx\wedge \xi }, \label{eq:RT}
\end{equation}
\begin{equation}
\widehat{{\bf T}}_\xi \widehat{{\bf R}}_x=\widehat{{\bf R}}_{x+\xi
/2}e^{-i2\pi Nx\wedge \xi }, \label{eq:TR}
\end{equation}
\begin{equation}
\widehat{{\bf R}}_{x_1}\widehat{{\bf R}}_{x_2}=\widehat{{\bf T}}%
_{2(x_2-x_1)}e^{i2\pi N(2x_1\wedge x_2)}. \label{eq:RR}
\end{equation}
We then have the same cocycle properties as in the plane: (\ref{eq:tt})-(\ref
{eq:rr}). This is a consequence of the commutation of operator products with
projection (\ref{eq:abpt}) and will be of crucial importance when we derive
the properties of the center and chord representations on the torus. Note
that the characterization of the chords $\xi $ by integers and the centers $%
x $ by half-integers is respected by the group of operations above.
Another property which results from the last cocycle relation (\ref{eq:RR})
is
\begin{equation}
\widehat{{\bf R}}_x\widehat{{\bf R}}_x=\widehat{{\bf 1}}_N^\chi ,
\end{equation}
in accordance with classical reflections. This means that
\begin{equation}
\widehat{{\bf R}}_x^{-1}=\widehat{{\bf R}}_x^{\dag }=\widehat{{\bf R}}_x,
\end{equation}
that is, reflection operators on the torus are unitary and Hermitian.
It is important at this stage to examine the trace of these operators. Using
(\ref{eq:TQQ}) and (\ref{eq:114}), we have:
\begin{equation}
{\bf Tr}(\widehat{{\bf T}}_\xi )=Ne^{i\frac{2\pi }N(\frac{rs}2+r\chi
_p-s\chi _p)}\delta _{r,0}^{(N)}\delta _{s,0}^{(N)}\equiv N\;e^{i\frac{2\pi }%
N(\frac{rs}2+r\chi _q-s\chi _p)}\;\delta _\xi ^{(N)}. \label{eq:trT}
\end{equation}
For the trace of the reflection operators, we recall (\ref{eq:RQQ}), so
\begin{equation}
<{\bf q}_n,N|\widehat{{\bf R}}_x|{\bf q}_n,N>=\delta _{n,2b-n}^{(N)}e^{i%
\frac{2\pi }Na(2b-2n)}\ , \label{eq:Rnn}
\end{equation}
which is different from zero only if
\begin{equation}
n=b\qquad \mbox{mod}\left( \frac N2\right) . \label{eq:nb}
\end{equation}
However, if $b$ is half-integer and $N$ is even, for example, there would be
no $n$ such that (\ref{eq:nb}) is satisfied. In general we can have up to 2
solutions of (\ref{eq:nb}) for $n\in \left[ 0,N-1\right] $, but they can
have different phase contributions in (\ref{eq:Rnn}). A careful inspection
leads to:
\begin{eqnarray}
{\bf Tr}(\widehat{{\bf R}}_x) &=&f_N(x)=\frac
12(1+(-1)^{2a}+(-1)^{2b}+(-1)^{2a+2b+N})= \nonumber \label{eq:trRT1} \\
&=&\left\{
\begin{tabular}{ll}
$0$ & $\mbox{ if N is even and }a\mbox{ or }b\mbox{ semi-integers}$ \\
$2$ & $\mbox{ if N is even and }a\mbox{ and }b\mbox{ integers}$ \\
$1$ & $\mbox{if N is odd and }a\mbox{ or }b\mbox{ integers}$ \\
$-1$ & $\mbox{ if N is odd and }a\mbox{ and }b\mbox{ semi-integers}$%
\end{tabular}
\right. \label{eq:trRT}
\end{eqnarray}
The importance of this result for the following theory calls for some
intuitive explanation in terms of the reflections of the discrete periodic
lattice. As in the plane case, we can relate ${\bf Tr}(\widehat{{\bf R}}_x)$
to the number of fixed points of the corresponding classical map. Indeed,
for $N$ odd there is always a single fixed point, agreeing with the modulus
of (\ref{eq:trRT}). If $N$ is even, there will only be fixed points if $x$
is characterized by integer numbers ($a,b$), in which case there are two.
\subsection{Operators and their Symbols}
Once we have defined the reflection and translation operators, we can
decompose any operator as their linear combination. To construct the chord,
or translation representation of an operator, we express any operator as a
linear combination of translations. To have a complete basis, we need just $%
N^2$ operators, so that $r$ and $s$ run from $0$ to $N-1$. The chords $\xi
=(\frac rN,\frac sN)$ having this property are said to belong to the
fundamental domain. The other translation operators are obtained from these
through the symmetry properties; that is, the fundamental translations are
those which have chords smaller than one loop around any of the irreducible
circuits of the torus in a given direction. The chord representation of an
operator is defined as
\begin{equation}
{\bf A}(\xi )\equiv {\bf Tr}\left( \widehat{{\bf A}}\widehat{{\bf T}}_{-\xi
}\right) . \label{eq:Acor}
\end{equation}
From the symbol, we recover the operator:
\begin{equation}
\widehat{{\bf A}}=\frac 1N\sum_{r,s=0}^{N-1}{\bf A}(\xi )\widehat{{\bf T}}%
_\xi \equiv \frac 1N\sum_\xi {\bf A}(\xi )\widehat{{\bf T}}_\xi .
\label{eq:corA}
\end{equation}
Although, to recover the operator we only need the symbol defined in the
fundamental domain (i.e. $r,s$ in $[0,N-1]$), (\ref{eq:Acor}) can be used to
extend the definition of the symbol for $r$ and $s$ running among all
integer numbers. Of course, these will not be independent of the symbols in
the fundamental domain and, from the symmetry properties of $\widehat{{\bf T}%
}(\xi )$ (\ref{eq:TTsim}), we see that ${\bf A}(\xi )$ satisfies
\begin{equation}
{\bf A}(\xi +{\bf k})=(-1)^{sk_p+rk_q+k_pk_qN}e^{i2\pi (k_p\chi _q-k_q\chi
_p)}{\bf A}(\xi )=e^{-i2\pi N\left[ \left( \frac \xi 2+\frac \chi N\right)
\wedge {\bf k}+\frac 14{\bf k}\widetilde{{\frak J}}{\bf k}\right] }{\bf A}%
(\xi ), \label{eq:Acorsim}
\end{equation}
where ${\bf k}=(k_p,k_q)$ is a vector with integer components denoting
chords that perform respectively $k_p$ and $k_q$ loops around the
irreducible circuits of the torus. This is an important consequence of the
fact that the symmetry properties of the symbol of operators are the same as
those of the basis operators used to generate this symbol.
We now expand the operators in term of reflections; this is the center or
Weyl representation. It is important to recall that we must take values of $%
a $ and $b$ that belong to $[0,\frac{N-1}2]$, that is, only one quarter of
the torus is needed to define a complete set of reflection operators . The
values of $x=(\frac{a+\chi _p}N,\frac{b+\chi _q}N)$ generated by these
values of $a$ and $b$ define the Weyl phase space ,WPS, shown by the bold
face area in Fig.~\ref{fig.3.2}.
We define the center symbol of an operator $\widehat{{\bf A}}${\bf \ }such
that,
\begin{equation}
{\bf A}(x)\equiv {\bf Tr}\left( \widehat{{\bf A}}\widehat{{\bf R}}_x\right) .
\label{eq:Acen}
\end{equation}
From the symbol, we recover the operator through
\begin{equation}
\widehat{{\bf A}}=\frac 1N\sum_{a,b=0}^{\frac{N-1}2}\widehat{{\bf R}}_x{\bf A%
}(x)\equiv \frac 1N\sum_x\widehat{{\bf R}}_x{\bf A}(x). \label{eq:cenA}
\end{equation}
The symmetry properties of $\widehat{{\bf R}}_x$ (\ref{eq:symR}) imply
\begin{equation}
{\bf A}(x+\frac{{\bf k}}2)=(-1)^{bk_p+ak_q+k_pk_qN}{\bf A}(x),
\label{eq:Acensim}
\end{equation}
for any vector ${\bf k}=(k_p,k_q)$ with integer components. This result had
already been obtained by Hannay and Berry \cite{hanay} for the Wigner
function and we see here that it is general for any Weyl symbol on the torus.
As in the plane case, we derive some important properties of the
translations and Weyl symbols. Notice first that:
\begin{equation}
{\bf R}_x(\xi )={\bf Tr}(\widehat{{\bf R}}_x\widehat{{\bf T}}_{-\xi })={\bf %
Tr}(\widehat{{\bf R}}_{x+\xi /2})e^{i2\pi Nx\wedge \xi }=f_N(x+\xi
/2)e^{i2\pi Nx\wedge \xi }
\end{equation}
and
\begin{equation}
{\bf T}_\xi (x)={\bf Tr}(\widehat{{\bf T}}_\xi \widehat{{\bf R}}_x)={\bf Tr}(%
\widehat{{\bf R}}_{x+\xi /2})e^{-i2\pi Nx\wedge \xi }=f_N(x+\xi /2)e^{-i2\pi
Nx\wedge \xi }.
\end{equation}
The trace is now obtained as
\begin{eqnarray}
{\bf Tr}\left( \widehat{{\bf A}}\right) &=&{\bf A}(\xi =0) \label{eq:trAcor}
\\
&=&\sum_x{\bf A}(x)f_N(x)=\frac 12\sum_{a,b=0}^{N-\frac 12}{\bf A}(x).
\end{eqnarray}
In the last equality we use the fact that the Weyl symbols for the entire
torus are obtained from those of a quarter of it through the symmetry
relations (\ref{eq:Acensim}) and the definition of $f_N(x)$ (\ref{eq:trRT1}).
The representation of the identity on the torus Hilbert space ${\cal H}%
_N^\chi $ has now the form:
\begin{equation}
{\bf 1}_N^\chi (x)=f_N(x)\ \ \ \mbox{ and }\ \ \ {\bf 1}_N^\chi (\xi
)=N\;\delta _\xi ^{(N)}.
\end{equation}
Hermitian operators are associated to the observables of the system and, in
particular, the Hamiltonian generates the dynamics. Defined as $\widehat{%
{\bf H}}=\widehat{{\bf H}}^{\dag }$, we obtain
\begin{equation}
{\bf H}^{\dag }(\xi )=[{\bf H}(-\xi )]^{*}\ \ \ \mbox{and}\ \ \ {\bf H}%
^{\dag }(x)=[{\bf H}(x)]^{*}\ ,
\end{equation}
just as for the plane case \cite{ozrep}.
The role played in the plane case by the Fourier transform will be taken by
the finite Fourier transform, since it allows us to exchange chords and
centers as well as to change from center or chord to the position
representation. But there are some small differences due to the $f_N(x)$
factors peculiar to the torus. Thus, in the exchange of centers and chords
we have,
\begin{eqnarray}
{\bf A}(\xi ) &=&{\bf Tr}\left( \frac 1N\sum_x\widehat{{\bf R}}_x{\bf A}(x)%
\widehat{{\bf T}}_{-\xi }\right) =\frac 1N\sum_x{\bf A}(x){\bf Tr}(\widehat{%
{\bf R}}_{x+\xi /2})e^{i2\pi Nx\wedge \xi } \nonumber \\
&=&\frac 1N\sum_x{\bf A}(x)f_N(x+\xi /2)e^{i2\pi Nx\wedge \xi },
\end{eqnarray}
whereas
\begin{equation}
{\bf A}(x)=\frac 1N\sum_\xi {\bf A}(\xi )f_N(x+\xi /2)e^{-i2\pi Nx\wedge \xi
}.
\end{equation}
Using (\ref{eq:TQQ}) and (\ref{eq:Acorsim}) we obtain the position
representation of an operator $\widehat{{\bf A}}$%
\begin{eqnarray}
<{\bf q}_m,N|\widehat{{\bf A}}|{\bf q}_n,N> &=&\frac 1N\sum_\xi {\bf A}(\xi
)<{\bf q}_m,N|\widehat{{\bf T}}_\xi |{\bf q}_n,N> \nonumber \\
&=&\frac 1N\sum_{r,s=0}^{N-1}{\bf A}(\xi _{r,s})e^{-i\frac{2\pi }Nr(\frac{m+n%
}2+\chi _q)}\delta _{m,n+s}^{(N)}e^{i\frac{2\pi }N(\frac r2+\chi _p)(m-n-s)}
\nonumber \\
&=&\frac 1N\sum_{r=0}^{N-1}{\bf A}(\xi _{r,m-n})e^{-i\frac{2\pi }Nr(\frac{m+n%
}2+\chi _q)}. \label{eq:AQsig}
\end{eqnarray}
Note that in this last equation we are employing chords that may not belong
to the fundamental domain; that is $m-n$ may not belong to $[0,N-1]$.
However, the symbol for this chord is well defined through (\ref{eq:Acor}).
If we restrict ourselves to chords that belong to the fundamental domain, we
then have a supplementary $e^{i\frac{2\pi }N(\frac r2+\chi _p)(m-n)}$ phase
factor in the last sum. This kind of difficulty may appear in the following
formulae, but, by allowing the indices to run over all integer numbers, the
formulae become indeed much simpler, as is the case for (\ref{eq:AQsig}).
Using the position representation
\begin{equation}
\widehat{{\bf A}}=\sum_{m,n=0}^{N-1}|{\bf q}_m,N><{\bf q}_m,N|\widehat{{\bf A%
}}|{\bf q}_n,N><{\bf q}_n,N|,
\end{equation}
we retrieve the chord representation as
\begin{eqnarray}
{\bf A}(\xi ) &=&{\bf Tr}\left( \sum_{m,n=0}^{N-1}<{\bf q}_m,N|\widehat{{\bf %
A}}|{\bf q}_n,N>|{\bf q}_m,N><{\bf q}_n,N|\widehat{{\bf T}}_{-\xi }\right)
\nonumber \\
&=&\sum_{m,n=0}^{N-1}<{\bf q}_m,N|\widehat{{\bf A}}|{\bf q}_n,N>e^{-i\frac{%
2\pi }Nr(\frac{m+n}2+\chi _q)}\delta _{m,n+s}^{(N)}e^{i\frac{2\pi }N(\frac
r2+\chi _q)(m-n-s)} \nonumber \\
&=&\sum_{n=0}^{N-1}<{\bf q}_{n+s},N|\widehat{{\bf A}}|{\bf q}_n,N>e^{-i\frac{%
2\pi }Nr(n+\frac s2+\chi _q)}. \label{AcorQ}
\end{eqnarray}
Using (\ref{eq:RQQ}) and (\ref{eq:Acensim}) we exchange the coordinate and
the center representation:
\begin{eqnarray}
<{\bf q}_m,N|\widehat{{\bf A}}|{\bf q}_n,N> &=&\frac 1N\sum_x{\bf A}(x)<{\bf %
q}_m,N|\widehat{{\bf R}}_x|{\bf q}_n,N> \nonumber \\
&=&\frac 1N\sum_{a,b=0}^{\frac{N-1}2}{\bf A}(x_{a,b})e^{i\frac{2\pi }%
N2(b-n)(a+\chi _p)}\delta _{m,2b-n}^{(N)}e^{i\frac{2\pi }N2\chi _p(m-2b+n)}
\nonumber \\
&=&\frac 1N\sum_{a=0}^{\frac{N-1}2}{\bf A}(x_{a,\frac{m+n}2})e^{i\frac{2\pi }%
N(m-n)(a+\chi _p)}. \label{eq:AQx}
\end{eqnarray}
Note that in this last formula we are taking the center point $x$ that does
not belong to the fundamental domain (i.e. $\frac{m+n}2$ may not belong to $%
[0,\frac{N-1}2]$ ). We recover the center representation through
\begin{eqnarray}
{\bf A}(x) &=&{\bf Tr}\left( \sum_{m,n=0}^{N-1}<{\bf q}_m,N|\widehat{{\bf A}}%
|{\bf q}_n,N>|{\bf q}_m,N><{\bf q}_n,N|\widehat{{\bf R}}_x\right) \nonumber
\\
&=&\sum_{m,n=0}^{N-1}<{\bf q}_m,N|\widehat{{\bf A}}|{\bf q}_n,N>e^{i\frac{%
2\pi }N(m-n)(a+\chi _p)}\delta _{m,2b-n}^{(N)}e^{i\frac{2\pi }Na(2b-n-m)}
\nonumber \\
&=&\sum_{n=0}^{N-1}<{\bf q}_{2b-n},N|\widehat{{\bf A}}|{\bf q}_n,N>e^{i\frac{%
2\pi }N2(b-n)(a+\chi _p)}. \label{AxQ}
\end{eqnarray}
It would be possible to define the chord and the Weyl representations by
equations (\ref{AcorQ}) and (\ref{AxQ}) respectively. However, the
geometrical structure, the role of the translations and reflection operators
and the relation to the plane theory would then be relegated to curiosities.
\subsection{Symbols of the product of operators}
We now derive the product law of the symbols of the operators in these
representations. Let us start with the chord representation (\ref{eq:corA}).
For this purpose we use ,(\ref{eq:TT}) and (\ref{eq:trT}) to obtain
\begin{eqnarray}
{\bf AB}(\xi ) &=&{\bf Tr}\left( \widehat{{\bf A}}\widehat{{\bf B}}\widehat{%
{\bf T}}_{-\xi }\right) =(\frac 1N)^2\sum_{\xi _1,\xi _2}{\bf A}(\xi _1){\bf %
B}(\xi _2){\bf Tr}\left( \widehat{{\bf T}}_{\xi _1}\widehat{{\bf T}}_{\xi _2}%
\widehat{{\bf T}}_{-\xi }\right) \nonumber \\
&=&(\frac 1N)^2\sum_{\xi _1,\xi _2}{\bf A}(\xi _1){\bf B}(\xi _2)Ne^{i2\pi
ND_4(\xi _1,\xi _2,-\xi )}{\bf Tr}\left( \widehat{{\bf T}}_{\xi _1+\xi
_2-\xi }\right) \nonumber \\
&=&(\frac 1N)\sum_{\xi _1}{\bf A}(\xi _1){\bf B}(\xi -\xi _1)e^{i2\pi
ND_3(\xi _1,-\xi )}, \label{eq:ABcor}
\end{eqnarray}
where we allow chords $\xi _2=\xi -\xi _1$ not to be in the fundamental
domain. Let us now take the trace of the product; inserting (\ref{eq:trAcor}%
) in (\ref{eq:ABcor}) leads to
\begin{equation}
{\bf Tr}\left( {\bf AB}\right) =(\frac 1N)\sum_{\xi _1}{\bf A}(\xi _1){\bf B}%
(-\xi _1).
\end{equation}
The generalization of (\ref{eq:ABcor}) for the product of an arbitrary
number of operators is
\begin{eqnarray}
{\bf A}_n...{\bf A}_1(\xi ) &=&(\frac 1N)^{n-1}\sum_{\xi _1...\xi _{n-1}}%
{\bf A}_1(\xi _1)... \nonumber \\
&&\!{\bf A}_{n-1}(\xi _{n-1}){\bf A}_n(\xi -\sum_{i=1}^{n-1}\xi _i)\exp
\left[ -i2\pi ND_{n+1}(\xi _1,...,\xi _{n-1},-\xi )\right] . \label{prodtr}
\end{eqnarray}
Thus, the product rule for the chords is obtained from that in the plane by
simply substituting the integral in (\ref{eq:ancor}) by the corresponding
sum.
For the center symbol (\ref{eq:Acen}) the trace of the product is obtained
using (\ref{eq:RR}) and (\ref{eq:trT}):
\begin{eqnarray}
{\bf Tr}\left( {\bf AB}\right) &=&(\frac 1N)^2\sum_{x_1,x_2}{\bf A}(x_2){\bf %
B}(x_1){\bf Tr}\left( \widehat{{\bf R}}_{x_2}\widehat{{\bf R}}_{x_1}\right)
\nonumber \\
&=&(\frac 1N)^2\sum_{x_1,x_2}{\bf A}(x_2){\bf B}(x_1)e^{i2\pi N(2x_1\wedge
x_2)}{\bf Tr}\left( \widehat{{\bf T}}_{2(x_2-x_1)}\right) \nonumber \\
&=&(\frac 1N)\sum_{x_1}{\bf A}(x_1){\bf B}(x_1).
\end{eqnarray}
We will now derive the full product properties in the center representation (%
\ref{eq:cenA}); with the help of the group properties (\ref{eq:RR}), (\ref
{eq:RT}) and (\ref{eq:trRT}) we have
\begin{eqnarray}
{\bf AB}(x) &=&{\bf Tr}\left( \widehat{{\bf A}}\widehat{{\bf B}}\widehat{%
{\bf R}}_x\right) =(\frac 1N)^2\sum_{x_1,x_2}{\bf A}(x_2){\bf B}(x_1){\bf Tr}%
\left( \widehat{{\bf R}}_{x_2}\widehat{{\bf R}}_{x_1}\widehat{{\bf R}}%
_x\right) \nonumber \\
&=&(\frac 1N)^2\sum_{x_1,x_2}{\bf A}(x_2){\bf B}(x_1)e^{i2\pi N2(x_1\wedge
x_2+x_2\wedge x+x\wedge x_1)}{\bf Tr}\left( \widehat{{\bf R}}%
_{x+x_2-x_1}\right) \nonumber \\
&=&(\frac 1N)^2\sum_{x_1,x_2}{\bf A}(x_2){\bf B}(x_1)e^{i2\pi N\Delta
_3(x,x_1,x_2)}f_N(x+x_2-x_1), \label{eq:ABcen}
\end{eqnarray}
where the symplectic area of the triangle $\Delta _3(x,x_1,x_2)$ was defined
in section 3. Note that the sides of these triangles must be integer vectors
in this case, because the symmetry of each side about its center implies
that all the corners will be of the same type regardless of whether either $%
a $ or $b$ are integer or half-integer. The argument of the function $f_N$
defined in (\ref{eq:trRT}) can thus be any corner of the triangle as shown
in Fig.~\ref{fig.3.3}(a). We thus find that the reflection properties of the QPS lead to
a more complex product rule than for the plane (\ref{eq:aacen}).
\begin{figure}[htb]
\centerline {\epsfxsize=5in \epsffile{fig3-3.eps} }
\caption{ Two examples of polygons displaying the uniform nature of
the vertices of the polygon. (a) all vertices lie on half integers: $f_N=0$
or $-1$ for $N$ respectively even or odd (b) the vertices lie on integers: $%
f_N=2 $ or $1$ for $N$ respectively even or odd.
}
\label{fig.3.3}
\end{figure}
The generalization for the product of $2n$ operators is
\begin{eqnarray}
&&{\bf A}_{2n}...{\bf A}_1(x)=(\frac 1N)^{2n}\sum_{x_1...x_{2n}}{\bf A}%
_{2n}(x_{2n}) \nonumber \\
&&{\bf ...A}_1(x_1)e^{i2\pi N\Delta
_{2n+1}(x,x_{2n},...,x_1)}f_N(x+\sum_{j=1}^{2n}(-1)^jx_j), \label{3.77}
\end{eqnarray}
where again the argument of $f_N$ is any corner of the polygon whose centers
are $x,x_{2n},...,x_1$ (see an example in Fig.~\ref{fig.3.3}(b)). For an odd number of
operators we just choose $\widehat{{\bf A}}_1=\widehat{{\bf 1}}$, that is $%
{\bf A}_1(x_1)=f_N(x_1)$ in (\ref{3.77}).
The product laws are the main result of this section. In contrast with the
Weyl-like representation obtained by Galleti and Toledo Pisa \cite{galeti2},
we only need half the number of sums (including the implicit sums in the trace of their formula (21)). Kaperskovitz and Peev \cite{kaperpeev}
also have a Weyl-like representation, but only for the case of $N$ even.
They perform products of 2 operators and the product law that they obtain is
very similar to ours, although their result is not compatible with our geometrical interpretation, because we need half-integer vectors to completely describe the reflections of QPS. Most important is the fact that our formalism prescribes the product of an arbitrary even number of operators, just as for the plane, whereas previous results could only cope explicitly with the product of two operators at a time.
\subsection{Weyl representation in QPS}
If $N$ is odd, we can redefine the WPS so that it coincides with the QPS.
For this purpose we define $X=(\frac{\alpha +\chi _p}N,\frac{\beta +\chi _q}%
N)$ so that ,
\begin{eqnarray}
\alpha &=&a+j\frac N2\qquad \mbox{ where }j=\left\{
\mbox{\begin{tabular}{ll}
$0$ & $\quad \mbox{ if }a\mbox{ is integer}$ \\
$1$ & $\quad \mbox{ otherwise}$\end{tabular}
}\right. \\
\beta &=&b+k\frac N2\qquad \mbox{ where }k=\left\{
\mbox{\begin{tabular}{ll}
$0$ & $\quad \mbox{ if }b\mbox{ is integer}$ \\
$1$ & $\quad \mbox{ otherwise}$\end{tabular}
}\right.
\end{eqnarray}
We then have that $\alpha $ and $\beta $ are integers for the case were $N$
is odd. In other words for any $x$ there is a point $X$ such that
\begin{equation}
X=\frac 1N\left(
\begin{array}{c}
\alpha \\
\beta
\end{array}
\right) +\frac \chi N=x+\frac 12{\bf n} \label{xentero}
\end{equation}
with ${\bf n}$ an integer vectors. If $N$ is even, the $\alpha $ and $\beta $
will have the same character ( integer or half-integer) as $a$ and $b$, so
we cannot recover the QPS. In the rest of this section we will then restrict
ourselves to the case where $N$ is odd. The symmetry relation (\ref{eq:symR}%
) shows then that
\begin{equation}
\widehat{{\bf R}}_X=(-1)^{(2jb+2ja+jk)}\widehat{{\bf R}}_x,
\end{equation}
and with the use of (\ref{eq:trRT}) we have
\begin{equation}
{\bf Tr}(\widehat{{\bf R}}_X)=1.
\end{equation}
We now see that letting $a$ and $b$ run over the half-integers in $[0,\frac{%
N-1}2]$, we then have $\alpha $ and $\beta $ integers in $[0,N-1]$ and we
recover the QPS. For this space we will now have a new Weyl representation
\begin{equation}
{\bf A}(X)={\bf Tr}\left( \widehat{{\bf A}}\widehat{{\bf R}}_X\right) ,
\end{equation}
from which we recover the operator:
\begin{equation}
\widehat{{\bf A}}=\frac 1N\sum_{\alpha ,\beta =0}^{N-1}\widehat{{\bf R}}_X%
{\bf A}(X_{\alpha ,\beta })\equiv \frac 1N\sum_X\widehat{{\bf R}}_X{\bf A}%
(X).
\end{equation}
We will now examine the properties of the product in this representation:
\begin{eqnarray}
{\bf AB}(X) &=&{\bf Tr}\left( \widehat{{\bf A}}\widehat{{\bf B}}\widehat{%
{\bf R}}_{X_{\alpha ,\beta }}\right) =(\frac 1N)^2\sum_{X_1,X_2}{\bf A}(X_2)%
{\bf B}(X_1){\bf Tr}\left( \widehat{{\bf R}}_{X_2}\widehat{{\bf R}}_{X_1}%
\widehat{{\bf R}}_X\right) \nonumber \\
&&(\frac 1N)^2\sum_{X_1,X_2}{\bf A}(X_2){\bf B}(X_1)e^{i2\pi N2(X_1\wedge
X_2-X_2\wedge X+X\wedge X_1)}{\bf Tr}\left( \widehat{{\bf R}}%
_{X+X_2-X_1}\right) \nonumber \\
&&(\frac 1N)^2\sum_{X_1,X_2}{\bf A}(X_2){\bf B}(X_1)e^{i2\pi N\Delta
_3(X,X_1,X_2)}. \label{eq:ABinter}
\end{eqnarray}
This last expression is very similar to the general case described in the
previous section, but slightly simplified by the absence of the $f_N$ term,
in close analogy to the plane formalism. For the product of $2n$ operators
this generalizes to
\begin{equation}
{\bf A}_{2n}...{\bf A}_1(X)=(\frac 1N)^{2n}\sum_{X_1...X_{2n}}{\bf A}%
_{2n}(X_{2n}){\bf ...A}_1(X_1)e^{i2\pi N\Delta _{2n+1}(X,X_{2n,...},X_1)}.
\end{equation}
The absence of the $f_N$ factor in these simplified formulae may be
understood from the fact that a polygon whose centers all lie on an integer
lattice always has corners on the same lattice. Hence, $f_N$ is always unity
for all corners if $N$ is odd.
In the same way, for $N$ odd, we can perform a transformation, similar to (%
\ref{xentero}), to a set of chords $\tilde{\xi}$ that are even multiples of $%
\frac 1N$. This set of chords will be complete if we now allow them to
perform up to two loops around the circuits of the torus. This scheme can be
generalized to perform quantization on centers or chords that are multiples
of $\frac \phi N$ only if $2\phi $ and $N$ are coprime numbers. Then, the
chord (or center) will be supported by a lattice of spacing $\frac \phi N$
and length $\phi $. This transformation will have importance for cat maps
and will be studied in more details in reference \cite{loxocat}.
\subsection{Relation between symbols}
There are many ways to represent a given operator $\widehat{A}$ that acts
on the Banach space of the plane ${\cal H}_{{\Bbb R}}$. Among the different
representations, the center and chord symbols are of special interest in
this work. Projecting the operator $\widehat{A}$ onto the torus Hilbert
space ${\cal H}_N^\chi $ through $\widehat{{\bf A}}=\widehat{{\bf 1}}_N^\chi
\widehat{A}\widehat{{\bf 1}}_N^\chi $, it can be represented in terms of
torus translations or reflections. We shall now show how the symbols on the
torus can be obtained from their counterparts on the plane.
Starting with the chord representation, we calculate the torus symbol at
points $\xi =\left( \frac rN,\frac sN\right) $. From the fact that $\widehat{%
{\bf T}}_{-\xi }$ and $\widehat{{\bf 1}}_N^\chi $ commute, we have
\begin{equation}
{\bf A}(\xi )={\bf Tr}\left( \widehat{{\bf 1}}_N^\chi \widehat{A}\widehat{%
{\bf 1}}_N^\chi \widehat{{\bf T}}_{-\xi }\right) ={\bf Tr}\left( \widehat{%
{\bf 1}}_N^\chi \widehat{A}\widehat{T}_{-\xi }\widehat{{\bf 1}}_N^\chi
\right) ={\bf Tr}\left( \widehat{A}\widehat{T}_{-\xi }\widehat{{\bf 1}}%
_N^\chi \right) .
\end{equation}
Then, we express the operator $\widehat{A}$ in terms of translations (\ref
{eq:acor}) and use the group properties of the translation operators (\ref
{eq:tt}) to obtain
\begin{equation}
{\bf A}(\xi )=\int \frac{d\xi _1}{(2\pi \hbar )}\ A(\xi _1)e^{\frac i{2\hbar
}\xi _1\wedge \xi }{\bf Tr}\left( \widehat{T}_{\xi _1-\xi }\widehat{{\bf 1}}%
_N^\chi \right) .
\end{equation}
We now use the projection properties of the translation operators on the
torus (\ref{projt}), so that
\begin{equation}
{\bf A}(\xi )=\int \frac{d\xi _1}{(2\pi \hbar )}\ A(\xi _1)e^{\frac i{2\hbar
}\xi _1\wedge \xi }{\bf Tr}\left( \widehat{{\bf T}}_{\xi _1-\xi }\right)
\left\langle \delta \left( \xi _1-\xi -\frac{{\bf k}}N\right) \right\rangle
_{{\bf k}}.
\end{equation}
Performing the integral, with the help of the trace properties (\ref{eq:trT}%
), we obtain
\begin{eqnarray}
{\bf A}(\xi ) &=&\left\langle (-1)^{sk_p+rk_q+k_pk_qN}e^{i2\pi (k_p\chi
_q-k_q\chi _p)}A\left( \frac rN+k_p,\frac sN+k_q\right) \right\rangle _{{\bf %
k}}= \nonumber \\
&&\left\langle e^{i2\pi N\left[ \left( \frac \xi 2-\frac \chi N\right)
\wedge {\bf k+}\frac 14{\bf k}\widetilde{{\frak J}}{\bf k}\right] }A\left(
\xi +{\bf k}\right) \right\rangle _{{\bf k}}, \label{eq:Acorprom}
\end{eqnarray}
where the $(2L)$-dimensional vectors ${\bf k}$ have integer components. Note
that we have to perform a phase weighted average on equivalent points to
obtain the symbol on the torus. This is similar to the way Hannay and Berry
quantize the cat map\cite{hanay} in the coordinate representation.
We now proceed in a similar manner to derive the symbols in the center
representation at the points $x=(\frac{a+\chi _p}N,\frac{b+\chi _q}N)$.
Using the commutation of $\widehat{{\bf R}}_x$ with $\widehat{{\bf 1}}%
_N^\chi $ and (\ref{eq:acor}), we have
\begin{equation}
{\bf A}(x)={\bf Tr}\left( \widehat{{\bf 1}}_N^\chi \widehat{A}\widehat{{\bf R%
}}_x\right) ={\bf Tr}\left( \widehat{{\bf 1}}_N^\chi \int \frac{dy}{(\pi
\hbar )}\ A(y)\hat{R}_y\widehat{R}_x\widehat{{\bf 1}}_N^\chi \right) ,
\end{equation}
which combined with the cocycle properties (\ref{eq:rr}), becomes
\begin{equation}
{\bf A}(x)={\bf Tr}\left( \widehat{{\bf 1}}_N^\chi \int \frac{dy}{(\pi \hbar
)}\ A(y)e^{\frac i\hbar 2y\wedge x}\widehat{T}_{2(x-y)}\widehat{{\bf 1}}%
_N^\chi \right) .
\end{equation}
Projecting the translations on the torus (\ref{projt}), we have
\begin{equation}
{\bf A}(x)=\int \frac{dy}{(\pi \hbar )}\ A(y)e^{\frac i\hbar 2y\wedge x}{\bf %
Tr}\left( \widehat{{\bf T}}_{2(x-y)}\right) \left\langle \delta \left(
2(x-y)-\frac{{\bf k}}N\right) \right\rangle _{{\bf k}},
\end{equation}
so that performing the integral we obtain,
\begin{eqnarray}
{\bf A}(x) &=&\left\langle e^{i2\pi (ak_q-bk_p+\frac N2k_qk_p)}A\left( \frac{%
a+\chi _p}N+\frac{k_p}2,\frac{b+\chi _q}N+\frac{k_q}2\right) \right\rangle _{%
{\bf k}} = \nonumber \\
& &\left\langle e^{i2\pi N(\left[ \left( x-\frac \chi N\right) \wedge {\bf k+%
}\frac 14{\bf k}\widetilde{{\frak J}}{\bf k}\right] )}A\left( x+\frac{{\bf k}%
}2\right) \right\rangle _{{\bf k}}, \label{eq:Axplator}
\end{eqnarray}
with the help of the trace properties (\ref{eq:trT}).
We again have a phase weighted average on equivalent points, this is a
general feature of any representation of projected operators ; only the
phase will depend on the specific representation we are taking. An important
feature of the center representation is that the phases do not have any
dependence on the $\chi $ parameters of the quantization; this is best seen
in (\ref{eq:Axplator}). Note also that comparing (\ref{eq:Acorprom}) and (\ref
{eq:Axplator}) with (\ref{eq:Acorsim}) and (\ref{eq:Acensim}) respectively,
the phases are a consequence of the periodicity conditions of the
symbols. It is important to note that if $\widehat{A}$ and $\widehat{{\bf 1}}%
_N^\chi $ commute the restriction of $\widehat{A}$ on the Hilbert space $%
{\cal H}_N^\chi $ denotes an automorphism. Indeed, the commutation of $%
\widehat{A}$ and $\widehat{{\bf 1}}_N^\chi $ implies that the symbols $A(x)$
and $A(\xi )$ are periodic functions. Otherwise the average defined in (\ref
{eq:Acorprom}) and (\ref{eq:Axplator}) may not exist, it may happen that the
projected operator $\widehat{{\bf A}}=\widehat{{\bf 1}}_N^\chi \widehat{A}%
\widehat{{\bf 1}}_N^\chi =0.$
\subsection{Symplectic invariance}
At the end of section 3, we remarked that the center and the chord
representations in the plane are invariant with respect to the quantum
equivalents of linear canonical transformations, or symplectic
transformations $x^{\prime }=Mx.$ The transformations for which the
symplectic matrix $M$ is made up of integers are known colloquially as {\it %
cat maps}. These have the property that they leave invariant the unit torus.
Because of the commutation of operator products with projection from the
plane to the torus, the effect of a similarity transformation $\widehat{{\bf %
A}}\rightarrow \widehat{{\bf U}}_M\widehat{{\bf A}}\widehat{{\bf U}}_M^{-1}$
performed by a quantized cat map on any operator defined on the torus will
be purely classical in the center or the chord representations:
\begin{equation}
{\bf A}(x)\rightarrow {\bf A}(Mx)\quad \mbox{and\quad }{\bf A}(\xi
)\rightarrow {\bf A}(M\xi ) . \label{cat1}
\end{equation}
Evidently, the matrix $M=1$ is a cat map; the product of cat maps is also a
cat map, as is the inverse of a cat map. It follows that the set of all cat
maps forms a subgroup of the symplectic transformations, which we will refer
to as the {\it feline group}. Likewise, relations (\ref{cat1}) indicate the
feline invariance of the chord and center representations.
In a companion paper \cite{loxocat} we use the chord and center
representations to study the properties of quantum cat maps of more than one
degree of freedom. This extends previous work by Hannay and Berry \cite
{hanay} and Keating \cite{keat1}, \cite{keat2} on two-dimensional cat maps.
For completeness, we will just note that the symbol corresponding to $%
\widehat{{\bf U}}_M$ is
\begin{equation}
{\bf U}_M(X)=\frac 1{\sqrt{N^L}}\exp \left( \frac i\hbar XBX\right) ,
\label{catx}
\end{equation}
whereas the chord symbol is
\begin{equation}
{\bf U}_M(\tilde{\xi})=\frac 1{\sqrt{N^L}}\exp \left( -\frac i\hbar \tilde{%
\xi}\beta \tilde{\xi}\right) . \label{catcor}
\end{equation}
For quantization performed on $\chi =0$, with $N$ an odd integer, $B$ and $%
\beta $ are integer symmetric matrices and the chords $\tilde{\xi}=\frac
2N(r,s)$ (we perform quantization on a set of chords that are multiple of $%
\frac 2N$). The symmetric matrices $B$ and $\beta $ in the above quadratic
forms define the Cayley parametrization of the symplectic matrix $M:$%
\begin{equation}
M=\frac{1-{\frak J}B}{1+{\frak J}B}=\frac{1+{\frak J}\beta }{1-{\frak J}%
\beta }.
\end{equation}
In general, The Cayley matrices for a cat map may not be integer, leading to
less transparent relation between the symplectic classical matrix $M$ and
the symmetric matrices $B$ and $\beta $ in (\ref{catx}) and (\ref{catcor}).
They depend on Gaussian sums, as presented in reference \cite{loxocat}.
\section{Hamiltonians on the Torus and Path Integrals}
\setcounter{equation}{0}
We will now treat Hamiltonian systems on the torus. Let us first recall that
the Poisson bracket relation that defines the symplectic product is the same
for the torus as in the case of the plane. Therefore the classical generating
function for canonical transformations are governed by the same composition
laws as defined in \cite{ozrep} for the plane case. The only difference is
that there would be different chords for a given center due to the periodic
boundary conditions that identify centers with half the period as that of
the whole torus.
We will then study dynamical systems with $L$ degrees of freedom for which
there is defined a Hamiltonian function that generates the dynamics through
Hamilton's equations and that is periodic in all its $2L$ variables. For $%
L=1 $ this kind of system has applications in solid state physics; it has
been used to model electron eigenstates in a one-dimensional solid with an
incommensurate modulation of the structure \cite{aubry} and in models of
Bloch electrons in a magnetic field \cite{harp}. It has also been shown \cite
{willk} that this model presents a critical behavior giving rise to
hierarchical structures in the solutions throughout the spectrum of the kind
known as a {\it Hofstadter butterfly} \cite{hofs} and to localization
transitions from extended to localized states.
The Fourier theorem ensures that a classical Hamiltonian that is periodic in
the plane can be written as
\begin{equation}
H(p,q)=\sum_{r,s=-\infty }^{+\infty }H_{r,s}e^{i2\pi (rp-sq)}.
\end{equation}
To quantize this Hamiltonian, different ways may be taken involving
different orderings. We choose the Weyl ordering, which is such that
\begin{equation}
\widehat{H}(\hat{p},\hat{q})=\sum_{r,s=-\infty }^{+\infty }H_{r,s}e^{i2\pi (r%
\hat{p}-s\hat{q})}.
\end{equation}
With the definition of the translation operators on the plane (\ref{eq:tcor}%
), we immediately see that this is equivalent to
\begin{equation}
\widehat{H}(\hat{p},\hat{q})=\sum_{r,s=-\infty }^{+\infty }H_{r,s}\widehat{T}%
_{\xi _{r,s}} \label{eq:hqt}
\end{equation}
where $\xi _{r,s}=(2\pi \hbar r,2\pi \hbar s)=(\frac rN,\frac sN)$. The
Hamiltonian is then a linear combination of translation operators that
leaves ${\cal H}_N^\chi $ invariant. If another ordering is chosen, there
will be corrections to (\ref{eq:hqt}) of the order of $\frac 1N$.
The quantal evolution of the system is determined by the propagator:
\begin{equation}
\widehat{U}_t=e^{\frac i\hbar t\widehat{H}}=\sum_{n=0}^\infty \frac
1{n!}(\frac i\hbar t\widehat{H})^n.
\end{equation}
This last relation implies that the propagator is a combination of products
of torus translations in the expansion (\ref{eq:hqt}). These form a cocycle,
as we already saw, so we can write
\begin{equation}
\widehat{U}_t=\sum_{r,s=-\infty }^{+\infty }U_{r,s}\widehat{T}_{\xi _{r,s}}.
\end{equation}
Thus, the evolution operator also leaves ${\cal H}_N^\chi $ invariant.
Written in this way we can see that the evolution operator and the
Hamiltonian in the periodic plane have their chord representation in terms
of torus translations only, in the form
\begin{equation}
U_t(\xi )=\sum_{r,s=-\infty }^{+\infty }U_{r,s}\delta (\xi_{r,s} -\xi).
\end{equation}
Let us now project this operator on ${\cal H}_N^\chi $ and follow the
evolution. We may first note that (\ref{eq:abpt}) and (\ref{eq:hqt}) allows
us to write:
\begin{eqnarray}
\widehat{{\bf U}}_t &=&\widehat{U}_t\widehat{{\bf 1}}_N^\chi
=\sum_{r,s=-\infty }^{+\infty }U_{r,s}\widehat{{\bf T}}_\xi
=\sum_{n=0}^\infty \frac 1{n!}(\frac i\hbar t\widehat{H})^n\widehat{{\bf 1}}%
_N^\chi \nonumber \\
&=&\sum_{n=0}^\infty \frac 1{n!}(\frac i\hbar t\widehat{H}\widehat{{\bf 1}}%
_N^\chi )^n=e^{\frac i\hbar t{\bf \hat{H}}} , \label{eq:UHT}
\end{eqnarray}
where ${\bf \hat{H}=}\widehat{{\bf 1}}_N^\chi \hat{H}\widehat{{\bf 1}}%
_N^\chi =\hat{H}\widehat{{\bf 1}}_N^\chi $ is the Hamiltonian acting in the
torus Hilbert space ${\cal H}_N^\chi .$
The unitarity evolution operators form a group, such that
\begin{equation}
\widehat{U}_t=(\widehat{U}_{\frac tM})^M.
\end{equation}
Projecting onto the torus and using (\ref{eq:abpt}) we obtain
\begin{eqnarray}
\widehat{{\bf U}}_t &=&\widehat{{\bf 1}}_N^\chi \widehat{U}_t\widehat{{\bf 1}%
}_N^\chi =\widehat{{\bf 1}}_N^\chi (\widehat{U}_{\frac tM})^M\widehat{{\bf 1}%
}_N^\chi \label{eq:prodevol1} \\
&=&(\widehat{{\bf U}}_{\frac tM})^M. \label{eq:prodevol2}
\end{eqnarray}
This last result is very important; the evolution and the projection commute.
We have two alternatives to obtain the center representation for (\ref
{eq:prodevol2}) , i.e. to work from the plane relations or to work directly
with the torus. First we note that (\ref{eq:UHT}) implies
\begin{equation}
\lim_{t\rightarrow 0}{\bf U}_t(x)=e^{\frac i\hbar t{\bf H}(x)}+0(t^2).
\end{equation}
The use of (\ref{eq:prodevol2}) and (\ref{3.77}) results in
\begin{eqnarray}
{\bf U}_t(x) &=&\lim_{M\rightarrow \infty }\left( \frac 1N\right)
^{2LM}\sum_{x_i=0}^{\frac{N-1}2}f_N(x+\sum_{j=1}^{2M}(-1)^jx_j) \nonumber \\
&&\exp \left\{ \frac i\hbar \left[ \,\Delta _{2M+1}(x,x_1,...,x_{2M})-\frac
t{2M}\sum_{i=1}^{2M}{\bf H(}x_i)\right] \right\} . \label{eq:utxT}
\end{eqnarray}
For the odd $N$ case we obtain the representation on the points $X$ of QPS,
\begin{equation}
{\bf U}_t(X)=\lim_{M\rightarrow \infty }\left( \frac 1N\right)
^{2LM}\sum_{X_i=0}^{N-1}\exp \left\{ \frac i\hbar \left[ \,\Delta
_{2M+1}(X,X_1,...,X_{2M})-\frac t{2M}\sum_{i=1}^{2M}{\bf H(}X_i)\right]
\right\} . \label{eq:utxTni}
\end{equation}
Notice that this expression for the projector relies on our original product rule for an arbitrary number of operators.
To take the projection, using (\ref{eq:prodevol1}), we can use the already
known result about the propagator in the center representation on the plane
\cite{ozrep}, obtained as a {\it path integral}
\begin{eqnarray}
U_t(x) &=&\lim_{M\rightarrow \infty }\int \frac{dx_1\cdots dx_{2M}}{(\pi
\hbar )^{2ML}}\ \exp \left\{ \frac i\hbar [\Delta _{2M+1}(x,x_1,\cdots
,x_{2M}))-\frac tM\sum_{n=1}^{2M}H(x_n)]\right\} \label{eq:utx2} \\
&=&\int_\gamma {\cal D}_\gamma \,e^{i/\hbar S_\gamma (x)}.
\end{eqnarray}
Here we can see that for the odd $N$ case the propagator (\ref{eq:utxTni})
is similar to (\ref{eq:utx2}) replacing the integral by the appropriate
sums. The phase of the integral in (\ref{eq:utx2}) coincides with the center
action $S_\gamma (x)$ for the polygonal path $\gamma $ with endpoints
centered on $x$ and whose $k^{\prime }$th side is centered on $x_k$. The
center variational principle ensures that this center action is stationary
for the classical trajectories centered on $x$. In Fig.~\ref{fig.1.8.}. we show two
possible paths, whose actions are compared by the center variational
principle.
\begin{figure}[h]
\vspace{7cm}
\caption{ Two possible paths, whose actions are compared by the center
variational principle.
}
\label{fig.1.8.}
\end{figure}
Now we project the symbol on the torus through (\ref{eq:Axplator}). We then
obtain
\begin{eqnarray}
{\bf U}_t(x) &=&\left( \frac 12\right) ^L\left\langle e^{i2\pi N(\left[
\left( x-\frac \chi N\right) \wedge {\bf k+}\frac 14{\bf k}\widetilde{{\frak %
J}}{\bf k}\right] )}\lim_{M\rightarrow \infty }\int \frac{dx_1\cdots dx_{2M}%
}{(\pi \hbar )^{2ML}}\right. \nonumber \\
&&\left. \exp \left\{ \frac i\hbar [\Delta _{2M+1}(x+\frac k2,x_1,\cdots
,x_{2M}))-\frac t{2M}\sum_{n=1}^{2M}H(x_n)]\right\} \right\rangle _{{\bf k}}
\nonumber \\
&=&\left( \frac 12\right) ^L\left\langle e^{i2\pi N(\left[ \left( x-\frac
\chi N\right) \wedge {\bf k+}\frac 14{\bf k}\widetilde{{\frak J}}{\bf k}%
\right] )}\int_\gamma d\gamma \,\;e^{\frac i\hbar S_\gamma (x+\frac{{\bf k}}%
2)}\right\rangle _{{\bf k}}. \label{eq:utxT2}
\end{eqnarray}
Although it is not immediately evident, (\ref{eq:utxT}) and (\ref{eq:utxT2})
are the same object; in (\ref{eq:utxT}), we first project on the torus and
then perform the evolution, while in (\ref{eq:utxT2}) we first evolve on the
plane and the projection on the torus is performed later. But, since the
projection (\ref{eq:prodevol2}) and evolution are commuting operations, (\ref
{eq:utxT}) and (\ref{eq:utxT2}) coincide. If we had defined the Weyl
transformation intrinsically in the torus, without projecting from the
plane, we could still derive a formula equivalent to (\ref{eq:utxT2}) with
the help of a Poisson transformation applied to (\ref{eq:utxT}).
To take the semiclassical limit, (\ref{eq:utxT2}) is the adequate
expression. Indeed, to apply this semiclassical limit we must evaluate the
integrals in (\ref{eq:utx2}) by the stationary phase approximation as in
\cite{ozrep}, so
\begin{equation}
U_t(x)_{SC}\sim \sum_j2^L|\det (1+{\cal M}_j)|^{-\frac 12}\ \exp \left\{
i\hbar ^{-1}S_{tj}(x)+i\mu _j\right\} , \label{eq:utxsc}
\end{equation}
where ${\cal M}$ is the symplectic matrix for the linearized transformation
between the neighborhood of the tips of the chord $\xi (x)$ generated by $%
S_t(x)$ as a center function.The index runs over all the contributing
classical orbits. In the case of a single orbit, the corresponding {\it %
Morse index} $\mu _j=0$. Hence on the torus we obtain
\begin{eqnarray}
{\bf U}_t(x)_{SC} &\sim &\left\langle e^{i2\pi N(\left[ \left( x-\frac \chi
N\right) \wedge {\bf k+}\frac 14{\bf k}\widetilde{{\frak J}}{\bf k}\right]
)}\right. \nonumber \\
&&\left. \sum_j2^L|\det (1+{\cal M}_j)|^{-\frac 12}\ \exp \left\{ i\hbar
^{-1}S_{tj}(x+\frac k2)+i\gamma _j\right\} \right\rangle _{{\bf k}}.
\label{eq:UTxsc}
\end{eqnarray}
The sum over $k$ is a sum over center points that are equivalent on the
torus because of the boundary conditions, but are different points on the
plane. To obtain the correct periodicity, the contribution of each term must
be summed with different phases. But for each point there are several
classical orbits whose chord is centered on it. The contribution of those
orbits are obtained in the $j$ sum. Then, for the semiclassical propagator
on the torus, we have a multiplicity of chords for any center point, due to
the boundary conditions.
\section{Conclusions}
\setcounter{equation}{0}
Our construction of the Weyl representation on the
torus naturally generates the conjugate chord representation.
This appears to be more useful on the torus than on the plane where it also
arises. The advantage of our derivation of the Weyl representation resides
in the clear geometrical interpretation of the operator basis in terms of
translations and reflections in QPS, so that the law for the symbol of the
product of operators acquires a simple form and generalizes to multiple products. It is important to note that
the parity of the number of states $N$ plays an important role and the
product law for $N$ odd is related to that in the plane case, by merely
replacing the integrals by the appropriate sums.
Although the geometric interpretation is valid for toral geometries, the
construction can be applied to any system whose Hilbert space has finite
dimension irrespective of the geometric structure of the underlying phase
space, except for its compactness. Indeed this operator basis and symbols
can be applied, for example, to spin systems or many-body fermionic systems
\cite{lipgal}. However, such a generalization destroys the intuitive
interpretation of the semiclassical limit.
By defining the operators on the torus as the projections of their analogues
on the plane, some important properties of the plane can then be used on the
torus. We exploit this fact for periodic Hamiltonian systems where we map
the continuous problem on a finite dimensional one. The path integral
formulation of Hamiltonian systems on the plane allows us to obtain that on
the torus, thus illuminating the semiclassical limit.
The symplectic invariance of the Weyl representation on the plane translates
to the torus as the Feline invariance; this fact will be used to study cat
maps of general dimension \cite{loxocat}.
{\it Acknowledgments: }We thanks A. Voros, M. Saraceno and R.O. Vallejos
for helpful discussions. We acknowledge financial support from Pronex-MCT
and A.M.F.R. also thanks support from CLAF-CNPq.
\clearpage
|
\section{Introduction}
A foam is a disordered collection of densely-packed polydisperse gas
bubbles in a relatively small volume of liquid
\cite{prudhommebook,djddaw,weaire84}.
Foams have a rich rheological behavior; they act like elastic solids
for small deformations but they flow like viscous liquids at large
applied shear stress \cite{kranikrev}. The stress is relaxed by
discrete rearrangement events that occur intermittently as the foam is
sheared. Three-dimensional foams are opaque, which makes it difficult
to observe these bubble movements directly. However, measurements
\cite{gopal,gopal2} by diffusing-wave spectroscopy of
three-dimensional foams subjected to a constant shear rate suggest
that the number of bubbles involved in the rearrangements is small, of
the order of four bubbles. Bubble rearrangements can be observed
directly by fluorescence microscopy in two-dimensional foams
found in insoluble
monolayers at the air-water interface. A study of shear in such
foams \cite{dennin}
also revealed no large-scale rearrangements.
While analytical theories for the response to applied steady shear may be
constructed for periodic foams, only simulation approaches are
possible for disordered foams. Kawasaki's \cite{okuzono} vertex model
was the first to incorporate dissipative dynamics. It applies to a
two-dimensional foam in the limit in which the area fraction of gas is
unity (a dry foam). Bubble edges are approximated by straight line
segments that meet at a vertex that represents a Plateau border. The
equations of motion for the vertices are solved by balancing
viscous dissipation due to shear flow within the borders by surface
tension forces. At low shear rates, the elastic energy of the foam,
which is associated with the total length of the bubble segments,
shows intermittent energy drops with a distribution of event rate vs.
energy release that follows a broad power law, consistent with
self-organized criticality. The rearrangements associated with the
largest events consist of cooperative motions of bubbles that extend
over much of the system.
Weaire and coworkers \cite{bolton0,bolton,hutzler} were the first to
develop a model appropriate to a disordered wet foam. The
model does not include dissipation, so it is quasi-static by
construction. Thus the system is allowed to relax to an equilibrium
configuration after each of a series of infinitesimal shear steps.
The size of rearrangements is measured by the number of changes in
nearest-neighbor contacts. For dry foams, the average event size is
small, inconsistent with a picture of self-organized criticality.
However, as the liquid content increases, the event-size distribution
broadens, with the largest events involving many bubbles. Although the
statistics are limited, this is consistent with a picture of
criticality at the point where the foam loses its rigidity.
The first model capable of treating wet, disordered foams at nonzero
shear rate was proposed by Durian \cite{durian1}. His model pictures
the foam as consisting of spherical bubbles that can overlap. Two
pairwise-additive interactions between neighboring bubbles are
considered, a harmonic repulsive force that mimics the effect of
bubble deformation and a force proportional to the velocity difference
between neighboring bubbles that accounts for the viscous drag. He
found \cite{durian2} that the probability density of energy drops
followed a power law, with a cutoff at very high energy
events. The
largest event observed consisted of only a few bubbles changing
neighbors. This is inconsistent with a picture of self-organized
criticality, although the effect of the liquid content on the topology
statistics was not
examined.
Most recently, Jiang et al.\cite{jiang} have employed a large-Q
Potts model to examine sheared foams. In this lattice model bubbles
are represented by domains of like spin, and the film boundaries are
the links between regions of different spins. Each spin merely acts
as a label for a particular bubble, and the surface energy arises only
at the boundaries where the spins differ. The evolution of the foam
is studied by Monte Carlo dynamics with a Hamiltonian consisting of
three terms: the coupling energy between neighboring spins at the
boundaries of the bubbles; an energy penalty for changes in the areas
of the bubbles, which inhibits coarsening of the foam; and a shear
term that biases the probability of a spin reassignment in the strain
direction. The spatial distribution of T1 events was examined and no
system-wide rearrangements were observed. Nevertheless, Jiang, et al.
found a power-law distribution of energy changes. They also found
that the number of events per unit strain displayed a strong shear
rate dependence, suggesting that a quasi-static limit does not exist.
These four simulation approaches thus offer conflicting pictures as to
(1) the existence of a quasistatic limit, (2) whether or not
rearrangement dynamics at low shear rates are a form of self-organized
criticality, and (3) whether or not the melting of foams with
increasing liquid content is a more usual form of criticality. One
possible reason for this disagreement is differences in the treatment of
dissipation, and hence in the treatment of the {\it dynamics} of the
rearrangements. In principle, the only accurate way in which to
include dissipation in a sheared foam is to solve for the Stokes flow
in the liquid films and Plateau borders. This approach has been
adopted by Li, Zhou and Pozrikidis\cite{pozrikidis}, but so far it has only
been applied to periodic foams. The statistics of rearrangement
events are fundamentally different in periodic and disordered foams;
in sheared periodic foams, all the bubbles rearrange simultaneously at
periodic intervals, while in a disordered foam, the rearrangements can
be localized and intermittent. Nonetheless, the Stokes-flow approach
is the only one that can be used as a benchmark for more simplified
models.
In order to gain a better understanding of the origin of the
discrepancies between the various models, as well as between the
models and experiments, we report here a systematic study of the
properties of a sheared foam using Durian's model. We begin by
reviewing his model and discussing our numerical implementation using
two different forms of dissipation. After confirming that there are
no significant system-size effects for dry samples, we examine
shear-rate dependence and establish the existence of a true
quasistatic limit for the distribution and rate of energy drops and
topology changes. This limit is shown to be independent of the dissipation
mechanism for foams of different gas fractions. Finally,
we examine dramatic changes in the behavior of these quantities as the
liquid content is tuned toward the melting point.
\section{Bubble model}
Durian's model\cite{durian1,durian2} is based on
the wet-foam limit, where the bubbles are spherical. The foam is
described entirely in terms of the bubble radii $\{ R_i \}$ and the
time-dependent positions of the bubble centers $\{ \vec{r}_i \}$. The
details of the microscopic interactions at the level of soap films and
vertices are subsumed into two pairwise additive interactions between
bubbles, which arise when the distance between bubble centers is less
than the sum of their radii. The first, a repulsion that originates
in the energy cost to distort bubbles, is modeled by the compression
of two springs in series with individual spring constants that scale
with the Laplace pressures $\sigma/R_i$, where $\sigma$ is the
liquid-gas surface tension and $R_i $ is the bubble radius. Bubbles
that do not overlap are assumed not to interact. The repulsive force
on bubble $i$ due to bubble $j$ is then
\begin{equation}
\vec{F}_{ij}^r = k_{ij} \left [(R_i + R_j) -
|\vec{r}_i - \vec{r}_j| \right ] \hat{r}_{ij}
\end{equation}
where $\hat{r}_{ij}$ is the unit vector pointing from the center of
bubble $j$ to the center of bubble $i$, and $k_{ij} = F_0/(R_i +R_j)$ is the
effective spring constant, with $F_0 \approx \sigma \langle R
\rangle$. The second interaction is the viscous dissipation due to
the flow of liquid in the films. It, too, is assumed to be pairwise
additive and is modeled by the simplest form of drag, where the force
is proportional to the velocity difference between overlapping
bubbles. The viscous force on bubble $i$ due to its neighbor $j$ is
\begin{equation}
\vec{F}_{ij}^v = -b(\vec{v}_i - \vec{v}_j),
\label{drag}
\end{equation}
where the constant $b$ is proportional to the viscosity of the liquid,
and is assumed to be the same for all bubble neighbors.
The net force on each bubble sums to zero, since inertial effects are
negligible in this system. Summing over those bubbles $j$ that
touch bubble $i$, the equation of motion for bubble $i$ is
\begin{equation}
\sum_j (\vec{v}_i - \vec{v}_j) = \frac{F_0}{b} \sum_j \left [
\frac{1}{|\vec{r}_i - \vec{r}_j|} - \frac{1}{R_i + R_j} \right ]
(\vec{r}_i - \vec{r}_j) + \frac{\vec{F}_i^a}{b},
\label{eq:vel}
\end{equation}
where $\vec{F}_i^a$ is an externally applied force, arising, for
instance, from interactions with moving walls.
Durian \cite{durian1,durian2} employed a further simplification of this model,
in
which the viscous dissipation is taken into account in a mean-field
manner by taking the velocity of each bubble relative to an average
linear shear profile. In this case, the total drag force on bubble $i$
due to all of its $N_{i}$ overlapping neighbors is
\begin{equation}
\vec{F}_{i}^v = -b N_{i} \left ( \vec{v}_i - \dot{\gamma}y_i \hat{x}\right ).
\label{mf}
\end{equation}
In the numerical simulations reported here we use both the mean-field
model of dissipation as well as the approximation represented by
Eq.~\ref{drag}, which we call the local dissipation model. In the
latter, at each integration time step the velocity of a bubble is
measured with respect to the average of the velocities of its $N_i$
overlapping neighbors, so that the total drag force on bubble $i$ is
\begin{equation}
\vec{F}_{i}^v = -b \left (N_{i} \vec{v}_i - \sum_{j={\text{
nn}}} \vec{v}_j \right )
\label{local}
\end{equation}
For very large $N_{i}$, this reduces to Eq.~\ref{mf}; otherwise, it
allows for fluctuations.
One aim of our study is to establish the sensitivity of the results to
the specific form of dissipation used, Eq.~\ref{mf} or
Eq.~\ref{local}.
In two dimensions, the area fraction of gas bubbles,
$\phi$, can be defined by the total bubble area $\sum\pi R_{i}^{2}$
per system area.
Because the bubbles are constrained to remain circular and their
interactions are approximated as pairwise-additive\cite{lacasse},
the model necessarily breaks down for very dry foams. In fact, bubble radii
can even be chosen so that $\phi$ exceeds one. In a real foam, of
course, this is prevented by the divergence of the osmotic pressure.
\section{Numerical Method}
All the results reported here are based on simulations of a
two-dimensional version of Durian's model. We use Eq.~\ref{eq:vel}
to study a two-dimensional foam periodic in the $x$--direction and
trapped between parallel plates in the $y$--direction. Bubbles that
touch the top and bottom plates are fixed to them, and the top plate
is moved at a constant velocity in the $x$--direction. (The system
can also be sheared with a constant force instead of a constant
velocity, but that case will not be discussed here.) Thus, bubbles
are divided into two categories --- ``boundary'' bubbles, which have
velocities that are determined by the motion of the plates, and
``interior'' bubbles, whose velocities must be determined from the
equations of motion.
The equation of motion Eq.~\ref{eq:vel} can be written in the form
\begin{equation}
\label{eq:Meom}
{\bf M}(\{{\bf r}\})\cdot\{{\bf v}\} = \{{\bf F}^r\}/b + \{{\bf
F}^a\}/b
\end{equation}
where $\{\bf v\}$ is a vector containing all the velocity components
of all of the bubbles, $\{v_0^x, v_0^y, v_1^x, v_1^y, \ldots\}$,
$\{{\bf F}^r\}$ is a vector of all of the repulsive bubble--bubble
forces, and $\{{\bf F}^a\}$ contains all the forces exerted by
the walls. The matrix $\bf M$ depends on the instantaneous positions
of the bubbles. The $2\times2$ block submatrix $M_{ij}$ is a unit
matrix $\bf 1$ if the distinct bubbles $i$ and $j$ overlap, and $\bf
0$ if
they do not overlap. On the diagonal, $M_{ii} = -{\bf 1} N_i$, where
$N_i$ is the number of overlapping neighbors of bubble $i$. Eq.
\ref{eq:Meom} is of the form ${\bf A}({\bf r}, t)\cdot (d{\bf r}/dt) =
f({\bf r}, t)$, which we solve for the bubble positions $\bf r$ with
the routine DDRIV3\cite{libs}. DDRIV3 has the ability to solve
differential equations in which the left hand side is multiplied by an
arbitrary time-dependent matrix. Furthermore, it allows all matrix
algebra to be performed by external routines, allowing us to take
advantage of the sparse nature of $\bf M$. We use the
SPARSKIT2\cite{libs} library for sparse matrix solutions.
The only relevant dynamical scale in this problem is set by the
characteristic relaxation time arising from the competing mechanisms
for elastic storage and viscous dissipation, $\tau_d= b\langle R
\rangle/F_0$. This is the characteristic time scale for the duration of bubble
rearrangements driven by a drop in total elastic energy. Without loss of
generality we set this to unity in the simulation. In these units,
the dimensionless shear rate $\dot \gamma$ is the capillary number.
To introduce polydispersity, the bubble radii are drawn at random from
a flat distribution of variable width; in all the results reported
here, the bubble radii vary from 0.2 to 1.8 times the average bubble
radius. We
note that the size distribution in experimental systems is closer to a
truncated Gaussian with the maximum size equal to twice the average
radius. The truncated Gaussian distribution arises naturally from the
coarsening process
\cite{glazierweaire,stavansrev}. We tested the sensitivity of our
results to the bubble distribution by doing one run with bubbles drawn
from a triangular distribution, and found that the shape of the
distribution had
no significant effect. Similarly, variation of the width of a triangular
distribution has been shown to have no influence on the linear
viscoelasticity \cite{durian2}. Note that it is important to include
polydispersity because a monodisperse system will crystallize under
shear, especially in two dimensions.
In all of our runs, the system is first equilibrated with all bubbles
treated as interior bubbles, and with a repulsive interaction between
the bubbles and the top and bottom plates so that bubbles cannot
penetrate the plates. The bubbles that touch the top and
bottom plates are then converted to boundary bubbles. The top plate
is moved at a constant velocity and data collection begins after any
initial transients die away. In addition to recording quantitative
measures of the system, we also run movies of the sheared foam in
order to observe visually how the flow changes as a function of shear
rate, area fraction and other parameters \cite{movies}.
\section{Quantities Measured}
Before showing results, we discuss the various quantities extracted
during a run. Under a small applied shear strain, bubbles in a real
foam distort; as the shear strain increases, the structure can become
unstable and they may thus rearrange their relative
positions. In the bubble model, the distortion of bubbles is measured
globally by the total elastic energy stored in all the springs connecting
overlapping bubbles:
\begin{equation}
E = \sum {1 \over 2}k_{ij} \left [(R_i + R_j) -
|\vec{r}_i - \vec{r}_j| \right ] ^{2}.
\label{Etot}
\end{equation}
Under steady shear, the elastic energy rises as bubbles
distort (overlap) and then drops as bubbles rearrange. Thus, the
total elastic energy fluctuates around some average value. The scale
of the energy is set by the elastic interaction and is of order $F_0
\langle R \rangle$ per bubble, where $ \langle R\rangle $ is the average
bubble radius.
Fig.~\ref{fig1}a shows a plot of the total elastic energy as a
function of strain for a system of 144 bubbles at area fraction
$\phi=1.0$ driven at a constant shear rate of $\dot \gamma =10^{-3}$.
Similar plots for stress vs strain are shown in Refs.
\cite{durian1,durian2}.
Note the precipitous energy drops, $\Delta E$, due to bubble
rearrangements. In the literature, these energy drops are often
referred to as avalanches. Since the term ``avalanche'' tends to
imply the existence of self-organized criticality, we employ the more
neutral but less elegant term ``energy drop.'' The time interval
between energy drops is much larger than the duration of a single
event. This is also illustrated in Fig.~\ref{fig1}b, which shows the
magnitude of energy drops that occur as the system is strained. ($\Delta
E$ is normalized by the average energy per bubble $E_b$, which has been
computed by averaging the elastic energy over the entire duration of a run
and dividing by the total number of bubbles in the system, $N_{bub}$.) These
recurring precipitous rearrangements represent the only
way for the foam to relax stress: there is no mechanism involving a
gradual energy release, as illustrated in Fig.~\ref{fig1}a. Note
that we compute only the total elastic energy of the system; because
events can be localized and intermittent, the elastic energy may be
dropping in one region of the sample and rising in other regions.
This would limit the size of the energy drop measured.
While useful for building intuition, the distribution of energy drops
does not yield direct information about bubble rearrangements.
Therefore, we also measure the number $N$ of bubbles that experience a
change in overlapping neighbors during an energy drop. We exclude
events in which two bubbles simply move apart or together; thus the
smallest event is $N=3$. A typical sequence of configurations before, during,
and after an event is shown in the first three frames of Fig.~\ref{edropav}.
In this energy
drop the magnitude of the drop and the number of bubbles that change
neighbors are close to the average. In the second and third frame of
the sequence, we have marked
the bubbles that changed neighbors since the beginning of the energy
drop (shown in the first frame). As the system is strained, more
bubbles change neighbors. For the particular energy drop chosen,
roughly one-sixth of the bubbles eventually change overlapping
neighbors. The fourth frame shows the final configuration of
bubbles (colored gray) superimposed on the initial configuration at
the start of the energy drop (colored black). Most of the bubble
motions that lead to this average-sized energy drop are rather subtle
shifts; there are no topological rearrangements. A large energy
drop, from the tail of the distribution, is shown in
Fig.~\ref{edropbig}. Again, the first three frames show the
configurations at the
beginning, middle and end of the drop, with the bubbles that change
overlapping neighbors marked in gray. The fourth frame shows the
extensive rearrangements that occur from the beginning to the end of
the drop. The configuration shown is the final one, and the short segments
are the tracks made by the centers of the bubbles during
the energy drop.
Typically, larger drops involve
larger numbers of bubbles. Fig.~\ref{fig1}c depicts $N$ during each
energy drop in the same run as in
Fig.~\ref{fig1}a and b. (Here, $N$ is normalized by the total number
of bubbles in the system, $N_{bub}$.) The correlation between energy drops
and
the
number of bubbles involved
is shown by a scatter plot of these quantities in Fig.~\ref{scat} for
a 900-bubble system strained from 0 to 10. We see that indeed there is a
strong correlation between these two measures of the size of an event.
Larger drops in energy involve larger numbers of bubbles and are
therefore spatially more extended. The correlation is particularly
good at the large-event end. There is more variability for midsize
and small events -- a large range of energy drops corresponds to the
same small number of rearranging bubbles, suggesting that typical
rearrangements involve only a few bubbles.
Besides counting statistics for energy drops and changes in number of
bubble overlaps, another direct measure of
bubble rearrangements is the rate of T1 events, i.e. of topology
changes of the first kind \cite{weaire84}. For a perfectly dry
two-dimensional foam consisting of thin films, these are said to occur
when a bubble edge shrinks to zero, such that a common vertex is shared by
four bubbles, two moving apart and two moving together. These events
were the only property used by Dennin and Knobler~\cite{dennin} to
characterize the response of their monolayer foam to shear because
they were unable to measure changes in the energy. While the time at
which a T1 event occurs is well defined in a dry foam, it is somewhat
ambiguous for a wet foam because there can be an exchange of nearest
neighbors without a common point of contact. Moreover, while the
number of bubbles involved in a T1 event is four by definition, large
clusters of bubbles can rearrange, with some of the interior bubbles
being involved in two or three T1 events simultaneously. It is then
much harder to assign an exact time to a T1 event.
To make contact with the monolayer experiments, we may define T1
events within the bubble model as follows. First we broaden the
definition of ``nearest neighbors'' to also include bubbles that do
not necessarily overlap, but that are nonetheless so close such that
$|\vec{r}_i - \vec{r}_j| < a(R_i + R_j)$, where $a>1$ is a suitably
chosen factor that may depend on $\phi$. We then say that a T1 event
begins when two nearest neighbors move apart, and we say that it ends
when a new nearest neighbor pair intrudes between them; the time at
which the event occurs is taken as the midpoint in this sequence.
This definition is illustrated in the time sequence of a T1 event
shown in Fig.~\ref{T1}. While the duration
of an actual T1 event in a dry foam is instantaneous, the duration
within the bubble model may vary greatly. Furthermore, the midpoint
in the sequence does not necessarily coincide with the exact moment
the switching occurs. In many instances it takes a long time after
two bubbles separate for the remaining pair to come into contact. To
compare with our other measures of rearrangement, we depict in Fig.~\ref{fig1}d
the number of T1 events as a function of strain for the same run
as in Figs.~\ref{fig1}a, b and c. There appears to be good correlation
between the largest energy drops and instances in which many T1 events
occur simultaneously. However, there are many more T1 events than
energy drops. This is because many T1 events can occur when a large
cluster of bubbles rearranges, and because our definition also includes
topology changes that cause an {\it increase} in the total elastic energy.
We can examine the consequences of our definition of a T1 event by
studying the distribution of the number of rearrangement events as a
function of their total duration in units of the strain. This is done for
both energy drops
and T1 events, as shown in Fig.~\ref{durdist}. The duration of an
energy drop is taken as the difference in strain between a decrease in the
elastic
energy and the next increase. It is evident from the duration
distribution for energy drops, Fig \ref{durdist}a, that most energy
drops occur over a relatively short strain scale. In units of time,
the longest events are comparable to a hundred times the
characteristic time scale in the problem ($\tau_d=1$ in our
simulations). We find a good correlation between the number of
bubbles that change overlapping neighbors and the duration of the
event; the more bubbles involved in the event, the longer it lasts.
The distribution for T1
events, shown in Fig \ref{durdist}b, has a qualitatively similar shape,
exhibiting a slightly more rapid decrease for both fast and slow
events. However, the scale on which T1 events occur is
an order of magnitude larger than the characteristic duration of the
energy drops. By examining the bubble motions we see that the largest
energy drops are associated with many T1 events,
but the difference in strain scales makes it difficult to demonstrate an
exact correlation between the number of overlap changes and the
number of T1's. In counting the T1 events, we include only events
that have a total strain duration of less than 2. Fig.~\ref{durdist}b
shows that we have included all the T1 events for this
run.
\section{Simulation Results}
For a given system size, strain rate, dissipation mechanism and gas
fraction, we now collect statistics on the
following measures of bubble dynamics: (1) The probability
distribution $P(\Delta E)$ for energy drops of size $\Delta E$;
(2) The probability distribution $P(N)$
for the number of bubbles $N$ that change overlapping neighbors during
a energy drop event; and (3) The event rates for both energy drops and
T1 events, ${S}(T1)$ and ${S}(\Delta E)$, both
defined as the number of events per bubble per unit strain.
\subsection{System Size}
We first address the important issue of the finite
size of the simulation sample. This is done for dry foams,
$\phi=1.0$, driven at a slow strain rate, $\dot \gamma=10^{-3}$. The
results for four system sizes, $N_{bub}=36$, 144, 324 and 900, are shown in
Fig.\ref{size}. In these runs, the systems were strained up to
80, 80, 31 and 10, respectively.
The top plot shows the energy drop distribution scaled by $E_b$, the
average energy per bubble. It shows that energy drops
vary greatly in size over the course of a single
run. The general features of this distribution have been reported
earlier \cite{durian2}. There is a power-law region with an exponent
of -0.7 that extends over several decades in $\Delta E/E_b$, followed
by a sharp cutoff that occurs above a characteristic event size. Such
a distribution has a well-defined average energy drop, which is near
the cutoff between 2$E_b$ and 3$E_b$ for the systems shown here. The
slight deviation from power-law behavior for small $\Delta E$ was
absent in the earlier simulations \cite{durian2}, which did not
exclude two-bubble events, and which had a different roundoff error. Also,
as seen earlier \cite{durian2}, the
two largest systems, with 324 and 900 bubbles, respectively,
have nearly identical
distributions. This has two important implications; namely, that the
sharp cutoff of the power-law distribution is not a finite-size
effect, and that the system does not exhibit self-organized
criticality.
The presence of a characteristic energy-drop size can be corroborated by
examining the number of bubbles that participate in rearrangements for the
same set of
runs, which is given in the middle plot, Fig. \ref{size}b. This quantity
has not been studied
previously
within the bubble model. We plot the probability distribution
$P(N)$ of the number of bubbles $N$ that change overlapping neighbors during a
rearrangement. The distribution decreases monotonically with a sharp
cutoff at the
large-event end. This indicates that most of the rearrangements are
local and involve only a few bubbles. Fig.~\ref{size}b shows that as the
system size
increases, the largest events represent a smaller fraction of the
total number of bubbles. Indeed, the tail of the distribution extends to
smaller and
smaller values of $N/N_{bub}$ with no signs of saturation as the system
size $N_{bub}$ increases, indicating diminishing finite size effects.
We next look at the system-size dependence of event rates,
$S(T1)$ and ${S}(\Delta E)$, for the number of T1
events and energy drops per bubble per unit strain. This is shown in
the bottom plot, Fig. \ref{size}c, for the same runs as in Figs.
\ref{size}a-b. We find that ${S}(\Delta E)$ decreases very
slightly with increasing system size, but saturates for the largest
systems. The results for ${S}(T1)$ show a stronger
system-size dependence, increasing slightly with $N_{bub}$. This could be
due to the fact that bubbles on the top and bottom boundaries of the
system are fixed, which lowers the number of possible T1 events per
bubble. As the system size grows, the boundary bubbles represent a
smaller fraction of the system so the event rate increases towards its
bulk value.
In short, all of our measurements at $\phi=1.0$ and $\dot
\gamma=10^{-3}$ indicate that the rearrangement events are localized
and that there is no self-organized criticality. This agrees with
observations of rearrangements in both monolayer and bulk foams.
\subsection{Shear Rate Dependence}
Now that size effects have been ruled out for dry foams, we may
examine the influence of shearing the sample at different rates.
Experiments by Gopal and Durian\cite{gopal2} on three-dimensional
foams show a marked change in the character of the flow with
increasing shear rate. At low shear rates, the flow is characterized
by intermittent, jerky rearrangement events occurring at a rate
proportional to the strain rate. As the shear rate increases, so that
the inverse shear rate becomes comparable to the duration of a
rearrangement event, the flow becomes smoother and laminar, with all
the bubbles gradually rearranging all the time. This was attributed
to a dominance of viscous forces over surface tension forces when the
strain rate exceeds the yield strain divided by the duration of a
rearrangement event.
In movies of our simulation runs, we also observe a crossover from
intermittent, jerky rearrangements to smooth laminar flow. Similar
smoothing has also been seen in stress vs. strain at increasing shear
rates for the mean-field version of bubble dynamics \cite{durian2}.
This raises the question of how the statistics of rearrangement events
change with shear rate. Specifically, how is the ``smoothing out''
of the flow reflected in the statistics at high
rates, and is there a quasistatic limit at low shear strain rates, in which
rearrangement behavior is independent of strain rate? Earlier numerical
studies by
Bolton and Weaire\cite{bolton} were restricted, by construction, to
the quasistatic limit. Okuzono and Kawasaki\cite{okuzono} examined
nonzero shear rates, but focused only on establishing the low
shear-rate limit. Recently, Jiang and coworkers found a strong
dependence of the T1 event rate on shear rate \cite{jiang}. They
found that the number of T1 events per bubble per strain,
${S}(T1)$, decreases sharply with strain rate with no evidence of a
quasi-static limit.
Our results for rearrangement behavior vs strain rate are collected in
Fig. \ref{shear} for a 144-bubble system at $\phi=1.0$. The top plot for
the probability distribution of energy drops indicates that there is
no gross change in $P(\Delta E)$ with shear rate, even though our
movies show a smoothing with less frequent energy drops. However,
there is some suppression of small energy drops with an accompanying
increase at large energy drops, as reflected in a somewhat smaller
power-law exponent and larger cutoff at high values of $\Delta
E/E_{b}$. It is not apparent from $P(\Delta E)$ vs $\Delta E/E_{b}$,
but we find that the average energy drop $\langle \Delta E \rangle$
and the average energy per bubble $E_{b}$ both increase with shear
rate, and that $\langle \Delta E \rangle$ increases more rapidly.
The reason why $E_{b}$ increases with shear rate is, of course, that
viscous forces become more important than elastic forces and lead to
increasing deformation (or in our model, overlaps) of bubbles. The
net result is that there are fewer, relatively larger, rearrangements
at high strain rates.
The tendency that small events are suppressed with increasing shear
rates is also borne out by the distribution of the number of bubbles
that change neighbors during an energy drop, as shown in
Fig.~\ref{shear}b. Note that unlike the previous curves, $P(N)$ is
plotted here on a linear scale. Two systematic trends emerge with
increasing $\dot{\gamma}$: there are relatively fewer small events,
i.e. $P(N)$ decreases significantly at small $N/N_{bub}$, and the tail extends
to slightly higher $N/N_{bub}$. For $\dot{\gamma} \geq 10^{-1}$ the
distribution is
fairly flat, suggesting that no one event size is dominant and there
are numerous large events of the order of the system size. This
suggests that at this shear rate the system no longer relaxes stress
by intermittent rearrangements, but by continuous flow, as confirmed by our
movies of the runs \cite{movies}. The trend
in $P(N)$ is seen in larger systems as well. For the 900-bubble system we
also find that as the shear rate increases from $10^{-5}$ to
$10^{-3}$, the distribution flattens and extends to higher values of
$N$. The average number of rearranging bonds increases
with shear rate, consistent with the picture of many bubbles in motion
as the system becomes more liquid-like. We cannot, however, probe the
system at very high shear rates. Data above a shear rate of about 1
cannot be trusted because of the nature of the model used. At high
rates of strain the viscous term dominates and the elastic forces are
not strong enough to prevent clumping of bubbles. This is actually an
artifact of the assumption that only overlapping bubbles interact
viscously; such clumping does not occur until much higher strain rates
in the mean-field version of dynamics. Another reason why we do not
study shear rates higher than unity is because we do not allow bubble
breakup under flow (recall that $\dot\gamma$ is the capillary number).
The gradual smoothing with increasing shear rate is most apparent in
Fig \ref{shear}c, where we see that the event rates of
T1 events and energy drops both decrease
with increasing strain rate. For the T1 events, the decrease is
slight, and is primarily due to the fact that the event duration
becomes even longer. The decrease is more
dramatic for the energy drop events. With increasing strain rate, the
average energy drop
increases and the rate of energy drops decreases.
Let us now re-examine the behavior of all quantities in
Fig.\ref{shear}, focusing on behavior at low shear strain rates.
Note that all quantities appear to approach a
reasonably well-defined ``quasistatic'' limit insensitive to the value
of $\dot\gamma$. We thus have the following picture. For small
$\dot{\gamma}$, the time between rearrangements is typically much
longer than the duration of a rearrangement, implying there is
adequate time for the system to relax stress. As the shear rate
increases, bubbles are constantly in motion and cannot fully rearrange
into local-minimum-energy configurations. Therefore, the viscous
interactions dominate, and the system flows like an ordinary liquid.
\subsection{Mean-Field vs Local Dissipation}
In the bubble model at higher strain rates, the behavior was seen to
depend on the form of dissipation: clumping for local dissipation,
Eq.~\ref{local}, as opposed to no clumping for mean-field dissipation,
Eq.~\ref{mf}. In this section we will investigate whether dissipation
affects the low-strain-rate behavior as well. If there truly exists a
quasi-static limit as $\dot{\gamma}\rightarrow 0$, as suggested by the
plots in the previous section, then the form of dissipation should
have no influence. This need not occur, since once a rearrangement
starts it proceeds with finite speed according to dynamics set by a
competition between surface tension and dissipation forces. For
example, it is conceivable that the mean-field dynamics might
discourage the mushrooming of a tiny shift in bubble position into a
large avalanche, whereas local dynamics might not. Another important
issue is that differences in mean-field vs local dissipation could be
relevant to true physical differences between bulk foams and Langmuir
monolayers at an air/water interface. For three-dimensional foams,
the shear is transmitted through the sample via bubble-bubble
interactions, so the dissipation might be better captured by the
local dissipation model. In contrast, for two-dimensional Langmuir
monolayer foams the subphase imposes shear on the monolayers, and the
dissipation might therefore be closer to that calculated with the
mean-field model.
To investigate the influence of mean-field vs local dynamics, we can
simply compare avalanche statistics. This is done in Fig. \ref{MFL}
for 144-bubble systems at four different area fractions, all sheared at
$\dot\gamma=10^{-3}$. The top plot shows results for
the energy-drop distribution, $P(\Delta E)$, with solid/dashed curves
for local/mean-field dissipation respectively. There is no
significant difference seen between the two choices of dissipative
dynamics. This is also true of the spatial extent of the
rearrangements, as seen in the middle plot for the probability
distribution $P(N)$ of rearranging bubbles. The bottom
plot for the rate of energy-drop and T1 events also shows little
significant difference between mean-field and local dynamics. The
only distinction is a slightly greater rate of T1 events in the
mean-field case. This reflects the difference in duration of T1
events within the two models; we find that T1 events tend to last
longer within the local dissipation model. Since we do not count T1
events that last longer than a strain of 2, we count fewer events within
the local model than the mean-field version. Thus, the differences in
$ S(T1)$ may simply be due to our method of counting T1
events. Taken together, the three plots in Fig.~\ref{MFL} encourage
us to believe that the rearrangement dynamics predicted by the model
are robust against details of the dissipation. They also provide
further evidence for the existence of a true quasi-static limit, where
the effect of strain rate is $\it{only}$ to set the rate of
rearrangements.
\subsection{Gas Area fraction}
Finally, we turn to the issue of how the elastic character of a foam
disappears with increasing liquid content, and possibility of critical
behavior at the melting transition. The principal signature of the
melting, or rigidity-loss,
transition is that the shear modulus $G=\lim_{t\rightarrow\infty}
\sigma(t)/\gamma$ vanishes and the foam can no longer support a
nonzero shear stress without flowing. In two-dimensional systems,
this happens at a critical gas fraction corresponding to that of
randomly packed disks, $\phi_c \approx 0.84$. This has been seen in
several different simulations, where the gas fraction was tuned to
within 0.05 of the transition\cite{bolton0,bolton,hutzler} and where
it was tuned through, and even below, the
transition\cite{durian1,durian2}. Other signatures of melting are
that the osmotic pressure vanishes as a
power-law\cite{durian1,durian2,lacasse} the coordination
number decreases towards about
4 as a power-law\cite{bolton0,bolton,hutzler,durian1,durian2,weaire}, and
that the time
scale for stress relaxation following an applied step-strain appears to
diverge \cite{durian1,durian2}.
Here we look for signs of melting in the
statistics of avalanches during slow, quasi-static flow. Within our
model, an increase in liquid content causes a decrease in the average
overlap between neighboring bubbles. This in turn produces a decrease
in the average elastic energy of the system, $E_{b}$ and sets the scale
for the average energy drop
$\langle\Delta E\rangle$ per rearrangement. It therefore should also
decrease at lower gas fractions $\phi$.
The energy drop and size statistics of rearrangement events for
increasingly wet foams were shown already in Fig.~\ref{MFL}, but were
discussed only in the context of mean-field vs local dissipative
dynamics. A clear trend emerges when we examine the $\phi$ dependence
specifically. In the top plot Fig.~\ref{MFL}a for $P(\Delta E)$, we
see that the power-law behavior for small events does not change, but
that the exponential cut-off moves towards larger values of $\Delta
E/E_b$ as $\phi\rightarrow\phi_{c}$. Though both $\langle\Delta
E\rangle$ and $E_{b}$ decrease towards zero, the latter evidently
vanishes more rapidly. This results in a broader distribution of
event sizes near the melting transition; as the system becomes more
liquid, large events are more prevalent.
The probability distribution $P(N)$ for the
numbers of bubbles involved in rearrangement events is shown in
Fig. \ref{MFL}b. It displays similar trends as a function of
$\phi$, but not as pronounced as in $P(\Delta E)$.
Namely, the power law for small $N$ is unaffected by $\phi$, but the
exponential cut-off moves towards slightly larger events as
$\phi\rightarrow\phi_{c}$. Thus, although the scale of energy drops
increases dramatically, the number of broken bonds only increases
marginally. Note,
however, that the largest events include almost all the bubbles in the
system; thus, the relatively weak dependence of $P(N)$ on $\phi$ could
be a finite-size effect in these $N_{bub}=144$ systems, as we will
show below.
The behavior of ${S}$, the number of energy drops and T1
events per bubble per strain, is shown in Fig.\ref{MFL}c. As the
system becomes wetter, there is no noticeable change in the event rate
${S}(\Delta E)$ for energy drops. In contrast, if our
definition of nearest neighbors only includes overlapping bubbles, we find
that ${S}(T1)$ decreases as $\phi$
decreases. This runs counter to expectations--bubbles in a wet foam
should have more freedom to move and rearrange because the energy
barrier between rearrangements is lower and the yield strain is
smaller. The apparent drop arises because the
bubble coordination number is much higher in a dry foam (roughly 6) than in a
wet foam (roughly 4). As a result there are more overlapping neighbors for
each bubble
in a dry foam, and more possibilities for the occurrence of T1 events.
In the wet foam, however, there are many T1 events that do not satisfy
the stringent starting or ending configurations because neighboring
bubbles do not overlap. It is therefore appropriate in wet foams to
modify the criterion for neighbors to $|{\bf r}_i - {\bf r}_j| < a(R_i
+ R_j)$, where the proximity coefficient $a$ is taken as $1/\phi$.
When T1's are computed with this definition, we find no significant
dependence on area fraction.
The fact that the power-law region of the energy drop distribution is
more extended at lower area fractions suggests the possibility of a
critical point as the close-packing density, $\phi_{c}$, is approached
from above. This would imply a pure power-law distribution $P(\Delta
E)$ for the energy drops at $\phi_{c}$, which would presumably be
accompanied by a growing correlation length, as well as the growing
relaxation time observed previously in Refs.~\cite{durian1,durian2}.
Note, however, that the distribution of the number of bubbles involved
in a rearrangement, $P(N)$, does not depend very strongly on $\phi$
for the 144-bubble systems of Fig. \ref{MFL}; furthermore, the
cut-off to power-law behavior is always present, no matter how closely
$\phi_{c}$ is
approached. This raises the question of whether finite system size
effects are more important at values of $\phi$ near $\phi_{c}$ (recall
from Fig.~\ref{size} that there were no significant system size
effects near $\phi=1$). To examine this, we have plotted the
dependence of $P(\Delta E)$, $P(N)$ and $ S$ on system size
in Fig.~\ref{size:85}. We indeed find a strong system size dependence
in $P(\Delta E)$ at $\phi=0.85$ just above the melting transition, with
no saturation at the largest size studied (900 bubbles). This is consistent
with the existence of a long correlation length.
The distribution of the number of bubbles per
energy drop, $P(N)$ also shows signs of
criticality. Recall from Fig.~\ref{size}b that at $\phi=1$, the tail of
$P(N)$ was cut off at smaller and smaller values of $N/N_{bub}$ with
increasing
system size at $\phi=1$. This was consistent with a short correlation
length, characteristic of localized rearrangement events. At
$\phi=0.85$, the behavior with increasing $N_{bub}$ is quite
different, as shown in Fig.~\ref{size:85}b. The distribution falls
off slightly more rapidly with $N/N_{bub}$ at larger system sizes
(probably because $\phi=0.85$ still lies above $\phi_{c}$), but the largest
events in the system still involve the same fraction $N/N_{bub}
\approx 0.75$ of bubbles, indicating a correlation length that is
comparable to the largest system size studied (30 bubble diameters
across).
The event rates for energy drops and T1 events for the
different system sizes at $\phi=0.85$ are shown in Fig.
\ref{size:85}c. The behavior is not markedly different from that
found for the drier foam. Recall, however, that we have adjusted our
definition of a T1 event by changing the proximity coefficient $a$
with area fraction, so little can be expected to be learned from this
measure.
\section{Discussion}
We have reported the results of several different measures of
rearrangement event dynamics in a sheared foam. A comparison of the
probability distribution of energy drops $P(\Delta E)$ with the
probabilty distribution of bubbles changing neighbors $P(N)$ shows
that the size of an energy drop correlates well with the number of
bubbles involved in a rearrangement (see Fig.~\ref{scat}). This is
valuable because the energy drop-distribution has been widely studied
theoretically, but is very
difficult to measure experimentally. The number of bubbles involved
in rearrangements, however, can be probed with multiple light
scattering techniques on three-dimensional foams\cite{gopal} and by
direct visualization of two-dimensional foams\cite{dennin}. A study
of the rate of occurrence of topological changes (T1 events) provides
a further link to experiments.
In general, our results agree with experiments on
three-dimensional and two-dimensional foams. Despite its simplicity,
the bubble model appears to capture the main qualitative features of a
sheared foam remarkably well. For example, we find that the size of
rearrangement events is typically small at
low shear rates and at area fractions not too close to $\phi_{c}$.
This is in accord with experiments of Gopal and Durian\cite{gopal},
and Dennin and Knobler\cite{dennin}, as well as simulation results of
Bolton and Weaire\cite{bolton} and Jiang and coworkers\cite{jiang}.
Our results do not agree with those of Okuzono and
Kawasaki\cite{okuzono}, however, who found power-law distributions of
rearrangement events at $\phi=1$ in two dimensions.
The largest discrepancies between our results and those of others lie
in the statistics of T1 events. We find that the number of T1 events
per bubble per unit strain is of order unity and is generally
insensitive to shear rate and gas area fraction. Kawasaki et al.
\cite {kawasaki} found similar results: ${S}(T1) = 0.5$ and
no dependence on shear
rate. In the Potts-model simulations \cite
{jiang}, however, ${S}(T1)$ is unity at $\dot{\gamma}=10^{- 3}$ but
falls to about 0.01
at $\dot{\gamma}=10^{- 1}$.
The monolayer experiments \cite{dennin} yielded values of $ S(T1)
\approx 0.15$, nearly an order of magnitude lower than predicted by
our simulations. Durian \cite{durian2} reported a number of rearrangement
events
per bubble per unit strain for simulations of a 900-bubble system at
$\dot{\gamma}=10^{- 5}$ that was comparable to the monolayer result,
but he measured the number of energy drops per bubble per unit
strain, $S(\Delta E)$, not the T1 event rate, $S(T1)$. Note that our
energy-drop event rate, $S(\Delta E)$, agrees well with Durian's
earlier result.
One might guess that the discrepancy between our measurement of
$ S(T1)$ and that of the monolayer experiment might lie in the
method of analysis used to count T1 events. Unlike the simulations,
in which the number of T1 events can be computed from an analysis of
bubble positions as a function of time, the number of T1's in the
monolayer studies was determined by repeated viewing of videotapes of
the experiments and counting of the events as the foam cells reach
their midpoint configuration. It seemed possible, then, that the
difference between the simulation and the experiment was the result of
a systematic undercounting of the number of the events. To check this
possibility, the number of T1's in a simulation run was determined by
observations of the animated bubble motions. The number of events
missed in this unautomated counting was only 2\% of the total.
We believe that the origin of the discrepancy between the T1 event
rates in the simulation and the monolayer experiment lies in the yield
strain. While the yield strain in the model system is less than 0.2,
which is consistent with that measured in three-dimensional foams,
that in the monolayer foams is closer to unity. Bubbles in monolayer
foams can therefore sustain very large deformations without inducing
rearrangements. The T1 event rate should be inversely proportional to
the yield strain. Thus, the ratio of $ S(T1)$ in the simulation
to $ S(T1)$ in the experiment should equal the ratio
of the yield strain in the experiment to the yield strain in the
simulation. This is exactly what we find.
One of our main results is that a quasistatic limit exists within
the bubble model. We find that the statistics of rearrangement events are
independent of shear rate at low shear rates. This agrees with the monolayer
experiments\cite{dennin}, which measured T1 event rates at two different shear
rates, $\dot
\gamma=0.003 s^{-1}$ and 0.11$s^{-1}$. Dennin and Knobler found no
noticeable difference in the T1 event rate, despite the fact that the shear
rates
studied differ by a factor of thirty. In addition, Gopal and Durian
found that the number of rearrangement events per bubble per second in a
three-dimensional foam is given by the event rate in the absence
of shear plus a term proportional
to the shear rate. In their case, the event rate was nonzero in the
absence of shear because of coarsening; we have neglected this effect
in our simulations. However, we do find that the rearrangement event
rate per unit time is simply proportional to the shear rate at low shear
rates.
Thus, experimental results in both two and three dimensions
contradict the simulation results of Jiang, et
al.\cite{jiang}, which find no quasistatic limit, but agree with our
findings.
The form of dissipation used in the
bubble model is a simple dynamic friction, which does not capture the
hydrodynamics of fluid flow in the plateau borders and films in a
realistic way. However, our results suggest that we may still be
capturing the correct behavior at low shear rates. We find that the
rearrangement event statistics are the same whether we use mean-field
or local dissipation at low shear rates.
This suggests that the statistics are determined by elastic effects rather
than viscous ones at low shear rates, and that the behavior in that limit
should be independent of the form of viscous dissipation used.
Finally, our results as a function of gas area fraction imply that
there may be a critical point at the melting transition, as the area
fraction approaches the random close-packing fraction from above.
Previous studies showed that both the shear modulus and yield stress vanish
as power laws at the melting
transition\cite{bolton,durian1}, and that the stress
relaxation time appears to diverge\cite{durian1}. Here, we have shown by
finite-size studies that there
is also a correlation length, characterizing the size of rearrangements,
which grows as one approaches the melting transition. We also find that the
distribution of energy drops appears to approach a pure power law in that
limit.
The existence of a critical point at the melting transition remains
to be tested experimentally. The vanishing of the shear modulus and
osmotic pressure at the transition has been measured by Mason and
Weitz\cite{mason} for monodisperse, disordered emulsions, and by Saint-Jalmes
and Durian for polydisperse gas-liquid foams\cite{arnaud}. However,
these small-amplitude-strain rheological measurements could not test
whether there is a
diverging length scale for rearrangements in a steadily sheared system
at the melting transition. On the other hand, Gopal and
Durian\cite{gopal} have measured the size of rearrangement events in
a gas-liquid foam, but only at packing fractions well above the melting
transition. At lower packing fractions close to the melting
transition, the liquid drains too quickly from the foam due to gravity
to permit such measurements. Experiments under microgravity
conditions should be able to resolve whether the melting transition is
indeed a critical point.
\acknowledgements We thank Narayanan Menon and Ian K. Ono for many
helpful discussions, and we thank Michael Dennin for performing the
visual analysis of the number of T1 events. This work was supported
by the National Science Foundation through grants CHE-9624090 (AJL),
CHE-9708472 (CMK), and DMR-9623567 (DJD), as well as by NASA through
grant NAG3-1419 (DJD).
|
\section{Introduction}
\indent
The current-voltage (I-V) relation of normal metal -superconductor
(NS) interfaces~\cite{btk}-\cite{lesovik} is strongly modified both by
wave interference phenomena (producing quasi-bound Andreev levels) and
ballistic transport at the interface. Several groups have observed
such novel superconducting phenomena at the NS interface between a
superconductor and semiconductor~\cite{kleins1}-\cite{poirier}. Based
on the success of Marsh et al.~\cite{marsh1}-\cite{marsh4} in
fabricating In and Sn `alloyed' contacts to a two-dimensional electron
gas (2DEG) formed at the AlGaAs/GaAs interface, we have studied
`alloyed' In contacts to the 2DEG. We find the mechanism for
producing highly transmissive NS contacts is In growth into the AlGaAs
`guided' along a preferred crystallographic direction. For an
AlGaAs/GaAs heterojunction with a [100] oriented surface, we find that
In growth into the AlGaAs occurs preferentially along the \{111\}
crystallographic planes. This unusual type of In growth into the
AlGaAs produces an `inverted pyramid' or `field emission' point
contact tip as shown in Fig.~\ref{cartoon1}(a). Similar microstructure
for metallic contacts to GaAs has been observed for both
AuGeNi~\cite{triangles} and Au~\cite{gold-GaAs} metallizations. This
guiding of In into the AlGaAs also allows the In to maintain its
superconducting properties. An AlInGaAs alloy, formed by diffusing In
into AlGaAs, would simply be a normal metal.
For such a crystallographically defined point contact metallization,
we find the closer one can grow the tip of the point contact to the
2DEG without contacting it, the higher the transmission coefficient of
an electron incident from the NS contact into the 2DEG. For such
nearly ballistic transport through the NS junction, a corresponding
excess current results~\cite{btk}. Growing the In down into direct
contact with the 2DEG, on the other hand, results in a
low-transmission normal metal - insulator - superconductor (NIS)
contact and its corresponding Giaever tunneling I-V
characteristic~\cite{btk}. We postulate that In in direct contact with
the 2DEG depletes the electrons around it, forcing the superconducting
electrons to tunnel through a large depletion layer near the contact,
shown schematically in Fig.~\ref{cartoon1}(b). Electron depletion
around a metallic contact to GaAs is commonly known as a Schottky
barrier. A similar mechanism for forming highly transmissive AuGeNi
contacts to GaAs was originally postulated by
Braslau~\cite{braslau}-\cite{woodall}. This mechanism for
forming highly transmissive tunneling type AuGeNi contacts to GaAs is
the reason some AlGaAs/GaAs heterojunction transistors can operate at
low temperatures with low contact resistance.
\begin{figure}
\centps{figs1a.eps}{45}
\centps{figs1b.eps}{45}
\caption{Schematic of the In profile obtained after annealing for (a)
a low temperature/short time anneal (Sample 1) and (b) a higher
temperature/longer time anneal (Sample 2). Penetration of In into
the AlGaAs layer is guided along a preferred crystallographic direction
in (a), forming point contacts to the 2DEG. The point contacts grow
together and penetrate the 2DEG in (b).}
\label{cartoon1}
\end{figure}
In this paper we correlate transmission electron microscope (TEM)
photographs of the superconducting In contacts to the resulting I-V
characteristics of the NS junctions. All the different $dI/dV$
characteristics shown in this paper are from nominally identical
samples, grown and prepared from the same GaAs wafer at the same
time. The only differences between the samples is in post process
contact annealing, and hence in the contact geometry.
Changes in contact geometry produce widely different $dI/dV$
characteristics. In point contacts grown near the 2DEG result in
ballistic transport of electrons through the NS contact and an excess
current. In in direct contact with the 2DEG produces lower
transmission contacts and Giaever tunneling. Since the GaAs
semiconductor forming the normal metal is also weakly localized, we
are able to observe weak localization corrections to Giaever
tunneling~\cite{vanwees1}. In In/GaAs junctions where some region of
the contact is transmissive, that portion of the contact will produce
a conductance drop around zero bias~\cite{marmorkos1}. The incoherent
addition of the conductance from different regions of the same
`contact' could therefore generate the `finite bias anomaly' seen in
Ref.~\cite{poirier}.
\section{Formation of the NS Contact}
\indent
A cross section of the unnannealed In/GaAs heterostructure is shown in
Fig.~\ref{heterostructure}(a). An undoped Al$_{0.3}$Ga$_{0.7}$As
spacer layer, followed by a Si doped Al$_{0.3}$Ga$_{0.7}$As layer and
a 50 $A^{o}$ protective Si doped GaAs layer, was grown on an undoped
(semi-insulating) (100) GaAs substrate. The resulting mobility of the
2DEG at liquid nitrogen temperature was about 125,000 $cm^{2}$/$V-s$.
The In contacts were deposited by thermal evaporation and liftoff. A
top view of the device is shown in Fig.~\ref{heterostructure}(b). Two
In pads, each of dimension 3.2mm$X$2.5mm and seperated by a nominal
gap of 4$\mu$, were deposited on top of the heterostructure using
thermal evaporation lift off. After annealing, we diffused In
through the AlGaAs barrier layer to contact the 2DEG.
The annealing temperature was varied between 500$^{o}$C and
600$^{o}$C. For temperatures of 450$^{o}$C or less the In failed
to contact the 2DEG, while annealing temperatures greater than
700$^{o}$C caused thermal deterioration of the interface. Large In
grains, 1-2 microns in diameter, grew on the contacts after annealing.
\begin{figure}
\centps{figs2a.eps}{45}
\centps{figs2b.eps}{45}
\caption{(a) Cross-section of the AlGaAs/GaAs heterostructure before
annealing the superconducting In contacts. (b) Top view of the
contact geometry.}
\label{heterostructure}
\end{figure}
A TEM micrograph of Sample 1, annealed for a relatively short duration
e.g. at 550 $^{o}C$ for 2 minutes, is shown in Fig.~\ref{tem1}(a).
Fig.~\ref{tem1}(a) shows that Indium starts growing into (100)
AlGaAs/GaAs preferentially along the $<$111$>$ directions. The growth
seems to be guided by the \{111\} crystallographic planes of GaAs
inclined at angles of about 55$^{o}$ to the surface. Therefore for
short annealing times we get `spikes' of Indium descending towards the
interface, forming an array of point contacts pictured in
Fig.~\ref{tem1}(a). The spikes are rather non-uniform in size and
irregular in their penetration depths, probably nucleating at defects
in the suface oxide of AlGaAs. In grain growth on the wafer surface
therefore has little effect on the final microstructure and geometry
of the NS contact. The net total conductance of the junction is
determined by the sum of the conductance of all the point contacts in
parallel. This growth mechanism is similar to that for alloyed
Au-Ge-Ni contacts with GaAs~\cite{woodall} in which most of the
conduction is through isolated Ge rich islands formed on the GaAs.
\begin{figure}
\begin{center}
\setlength{\epsfxsize}{7.0truecm}
\setlength{\epsfysize}{7.0truecm}
\epsfbox{image3-2.eps}
\setlength{\epsfxsize}{7.0truecm}
\setlength{\epsfysize}{7.0truecm}
\epsfbox{image2-2.eps}
\end{center}
\caption{TEM photographs of an In-AlGaAs/GaAs contact annealed (a) for
a short time (Sample 1) and (b) a longer time (Sample 2). Penetration
of In into the sample is guided by the \{111\} planes, forming the In
point contacts are clearly observable in (a). The point contacts
agglomerate and grow through the AlGaAs/GaAs interface in (b).}
\label{tem1}
\end{figure}
A TEM micrograph of Sample 2, annealed at 550$^{o}$C for 6 minutes and
then at 650$^{o}$C for 3 minutes, is shown in Fig.~\ref{tem1}(b). For
the longer annealing times and higher annealing temperatures used in
sample 2, more such In spikes grow from the deposited In
contact. Fig.~\ref{tem1}(b) shows these In spikes coalesce and
penetrate completely through the AlGaAs/GaAs interface to physically
touch the 2DEG, as depicted schematically in Fig.~\ref{cartoon1}(b).
The conductance characteristics of Samples 1 and 2 differ dramatically
as described in the next section.
A high magnification TEM photograph of one of the inverted pyramid
type In spikes in Sample 1 is shown in Fig.~\ref{tem2}. The (100) GaAs
surface is towards the top of Fig.~\ref{tem2}. The penetration of In
into the AlGaAs is clearly guided by \{111\} crystallographic
planes. One can see in the high resolution TEM picture that the In
indeed follows the \{111\} AlGaAs planes for several atoms. The
In-AlGaAs boundary then moves abruptly along the [010] direction for
1-2 atoms before continuing along the \{111\} planes. The detailed TEM
picture in Fig.~\ref{tem2} shows that while the \{111\} planes
strongly guide the growth of In into (100), the In does not exactly
follow those planes. However, the overall structure of the In
contacts still resembles a point contact.
\begin{figure}
\begin{center}
\setlength{\epsfxsize}{7.0truecm}
\setlength{\epsfysize}{7.0truecm}
\epsfbox{image1small.eps}
\end{center}
\caption{High magnification TEM photograph of a portion of a single In
point contact grown into a (100) AlGaAs surface. In growth into the
AlGaAs is strongly guided along the \{111\} planes.}
\label{tem2}
\end{figure}
\section{Current-Voltage Characteristics}
\indent
Hall measurements were made on the 2DEG at the interface at a
temperature of 400 mK. From the slope of transverse resistance
$R_{xy}$ we estimate the carrier density to be $1.8X10^{11}$
$cm^{-2}$. The mobility was 215,000 ${{cm}^2}/{V-s}$ which yielded a
mean free path $l$ of about 0.4 $\mu$. From the low-field
magnetoresistance, we estimate the phase breaking length $l_{\phi}$ to
be about 1.6 $\mu$m~\cite{chaud-thesis}, which is less than the 4
$\mu$m separation between the two In pads in
Fig.~\ref{heterostructure}(b). We conclude that the conductance of the
devices is essentially equivalent to two NS junctions in series
separated by a series resistor (the 2DEG).
\subsection{NS Junction with an excess current}
\indent
In Fig.~\ref{base12}(a) we show the conductance characteristics of
Sample 1 at a base temperature of 100 mK. Fig.~\ref{base12}(a) shows
an increase in conductance of about 10-12\% around a range of $\pm$
6mV. The 10-12\% excess conductance around zero voltage shows that
the junction behaves like a moderately transmissive
junction~\cite{btk}, with the majority of the diffused Indium spikes
forming transmissive interfaces with the 2DEG of the GaAs. The large
oscillations of the conductance in Fig.~\ref{base12}(a) are
reproducible with thermal recycling of the device. The oscillations
are therefore likely a consequence of electron wave interference due
to scattering from the fixed In point contacts to the 2DEG. The
parasitic resistance of the 2DEG in series with the two NS junctions
stretches the dI/dV versus V characteristics along the voltage axis,
and also suppresses any changes observed along the dI/dV axis. Series
resistance of the 2DEG explains why we observe only a 10-12\% excess
conductance instead of an excess conductance approaching 100\% in
Fig.~\ref{base12}(a).
\begin{figure}
\centps{figs5a.eps}{60}
\centps{figs5b.eps}{60}
\caption{Conductance vs the DC bias voltage across the two Indium pads
for (a) Sample 1 and (b) Sample 2. Sample 1 shows nearly 100\%
ballistic transport of electrons (after correcting for series
resistance), while Sample 2 shows Giaever tunneling (through a
tunnel barrier having transmission of the order $T \simeq 0.1$).}
\label{base12}
\end{figure}
We can crudely estimate the interface transmission in
Fig.~\ref{base12}(a) by comparing it to an ideal, ballistic NS
junction~\cite{btk}. A perfectly transmissive interface shows a 100\%
rise of conductance when the voltage bias satisfies $|eV| \leq
\Delta$, where $2\Delta$ superconducting energy
gap~\cite{btk}. Setting $2 \Delta =$ 6mV (for two NS junctions in
series), and using the BCS formula $2 \Delta = 3.5 k_B T_c$, gives a
critical temperature $T_c = 20$K. Since the actual critical
temperature of the contacts is about $3.4$K, we infer a 1mV drop
across the two NS interfaces and 5mV across the semiconductor. The
total conductance at the gap voltage can be read directly from
Fig.~\ref{base12}(a) as about 9mS, from which we infer a series
resistance of about 92.6$\Omega$ and a resistance of the two NS
interfaces in series of about 18.5$\Omega$ (when the voltage across
the NS interface is less than the energy gap). When the voltage across
the two NS interfaces is large, we can read directly off
Fig.~\ref{base12}(a) a conductance of about 7.75mS, or a total
resistance of about 129$\Omega$. Subtracting the series resistance, we
see the device resistance changes from 18.5$\Omega$ when $V_{\rm
interface} \leq \Delta$ to about 36.4$\Omega$ when $V_{\rm interface}
\gg \Delta$, roughly a 97\% increase in background conductance. Since
strong wave interference is present in Fig.~\ref{base12}(a), this
comparison should be regarded as giving an order of magnitude
estimate. However, after correcting for series resistance, this
estimate indicates the junction in Fig.~\ref{base12}(a) is a nearly
ballistic NS interface.
\subsection{Giaever-type Tunnel Junctions}
\indent
In Fig.~\ref{base12}(b) we show the differential conductance for
Sample 2 at a temperature of 400mK. The differential conductance is
suppressed around zero voltages, indicating Giaever tunneling and a
lower transmission contact. We can conclude from Figs.~\ref{cartoon1}
and \ref{base12} that In in intimate contact with the 2DEG forms a
relatively low transmission contact, whereas In nearby but not
directly in contact with the 2DEG forms a high transmission contact.
The temperature dependence of the differential conductance for Samples
1 and 2 is also consistent with a ballistic contact and a low
transmission contact, respectively. The differential conductance of
Sample 1 is relatively constant with temperature, while the
differential conductance of Sample 2 greatly decreases as the
temperature is lowered (indicating thermionic emission).
We have taken several NS contacts displaying Giaever tunneling at low
temperature and annealed them for longer times, attempting to obtain a
ballistic NS interface. In all cases further annealing simply makes
the Giaever tunneling characteristic more pronounced, indicating that
further annealing lowers the interface transmission. The room
temperature conductance of the sample improves with further annealing,
however, simply due to an increase of the effective contact
area. Since the room temperature conduction mechanism through the
contact of Sample 2 is thermionic emission, it simply scales with the
contact area. Further annealing improves the room temperature
conductance, but worsens the low temperature conductance. There is an
optimal annealing time where In grows down to almost reach the 2DEG,
but is not in physical contact with the 2DEG. Any further annealing
after this point degrades interface transmission.
\subsection{Effect of Weak Localization on Giaever Tunneling}
\indent
For an NS junction of length L, obeying the condition $l\ll L \leq
l_{\phi}$, electrons which initially failed to Andreev refelct from
the NS interface can backscatter again to the NS interface.
Therefore, weak localization inside the normal conductor gives the
electrons additional opportunities for Andreev reflection. The net
effect of Giaever tunneling at the NS junction combined with
weak localization inside the normal conductor is an enhancement of the
total Andreev reflection probability at the Fermi level, leading to an
additional conductance peak around zero bias
voltage~\cite{kleins1,vanwees1}.
Fig.~\ref{weakloc} shows the differential conductance for the NS
junction annealed at 550$^{o}$C for 3 minutes, which we call Sample
3. Sample 3 displays the conductance peak around zero bias voltage
first observed in Ref.~~\cite{kleins1} and explained in
Ref.~\cite{vanwees1}. Disorder assisted backscattering can cause a
zero bias conductance peak of magnitude up to 10\% of the background
conductance value. In our case the zero bias peak is about a 1.5\%
increase over the background conductance. The Giaever tunneling
feature is also spread over a large voltage range larger than 1mV.
Parasitic resistance from the 2DEG again explains the smaller zero
bias conductance peak and the spreading out of the dI/dV versus V
along the voltage axis. Since the height of the zero bias conductance
peak saturates by 300 mK Fig.~\ref{weakloc}(a), the peak is not the
precursor of a supercurrent between the two contacts. The
supercurrent should be negligibly small in any case, since the sample
satisfies $l\ll L \geq l_{\phi}$. We adequately filtered RF noise away
from the devices to observe supercurrents in other superconductor -
semiconductor samples with closer pad separation. We would also have
observed such a supercurrent if it were present in this sample.
Fig.~\ref{weakloc}(b) shows the magnetic field dependence of the
conductance for Sample 3. For a junction of length L and width W,
where $L,W > L_{\phi}$, the magnetic field required to destroy this
zero bias conductance peak is of the order
$B_{c}={\phi_{0}}/{{L_{\phi}}^2}$. In Sample 3 we observe $B_{c}
\simeq$ 80 Gauss. The calculated field is $B_{c} \simeq$ 20 Gauss,
nearly four times smaller than the observed value. It is possible that
the $L_{\phi}$ is overestimated, i.e. it may be around
$L_{\phi}\simeq$ 0.8 $\mu$m. Since these numbers are not precise data
fits based on any quantitative theory, it is comforting that we obtain
roughly the coherence length obtained by previous weak localization
measurements on the 2DEG. The shift in the background conductance
with the magnetic field in Fig.~\ref{weakloc}(b) is also due to the
parasitic magnetoresistance of the 2DEG.
\begin{figure}
\centps{figs6a.eps}{60}
\centps{figs6b.eps}{60}
\caption{ Conductance vs the DC bias voltage for Sample 3 for (a)
different temperatures and (b) different magnetic fields. A
conductance peak develops near zero bias voltage, a correction to
Giaever tunneling arising from weak localization inside the normal
metal.}
\label{weakloc}
\end{figure}
\subsection{Anamalous Weak Localization Corrections to Giaever
Tunneling}
\indent
On most samples where we observed weak localization corrections to
Giaever tunneling, we obtained conductance characteristics similar to
those in Fig.~\ref{weakloc}. However, Fig.~\ref{anomloc} shows the
conductance characteristics for an NS junction annealed at 500$^{o}$ C
for 2 minutes which we call Sample 4. Overall Sample 4 displays a
background of Giaever tunneling. A zero bias conductance peak (similar
to the one in Fig.~\ref{weakloc}) continues to develop for
temperatures down to about 800 mK in Fig.~\ref{anomloc}(a). At a
temperature of 650 mK in Fig.~\ref{anomloc}(a), however, a conductance
dip begins developing around zero bias. This dip in conductance around
zero bias, which is superposed on the broader conductance peak, is
nearly fully developed by 300 mK as shown in Fig.~\ref{anomloc}(a).
There is little change in the differential conductance between 300 mK
and 180 mK in Fig.~\ref{anomloc}(a). This anomalous dip feature
superimposed on the weak localization correction to Giaever tunneling
is reproducible on thermally cycling the NS junction back to room
temperature and again down to mK temperatures.
\begin{figure}
\centps{figs7a.eps}{60}
\centps{figs7b-2.eps}{60}
\caption{ Conductance vs the DC bias voltage for Sample 4 for (a)
different temperatures and (b) different magnetic fields. The additional
conductance dip which develops around zero bias voltage can be explained by an
inhomogeneous NS contact consisting of both high and low transmission regions. }
\label{anomloc}
\end{figure}
Marmorkos, Beenakker, and Jalabert~\cite{marmorkos1} have numerically
simulated the conductance of an NS junction in contact with a dirty
normal metal. For low transmission interfaces they numerically
observe, in Fig.~2 of Ref.~\cite{marmorkos1}, the zero bias
conductance peak associated with the weak localization corrections to
Giaever tunneling. However, for high transmission between the NS
interface and normal conductor, the numerical simulations of
Ref.~\cite{marmorkos1} reveal that the conductance peak changes into a
conductance dip around zero bias. Ref.~\cite{marmorkos1} therefore
shows that the same weak localization phenomena which causes a zero
bias conductance peak in low transmission contacts causes a zero bias
conductance dip for highly transmissive NS interfaces.
The numerical simulation in Ref.~\cite{marmorkos1} offers one possible
way to explain the conductance dip around zero bias we observe in
Sample 4. The overall conductance of Sample 4 displays Giaever
tunneling. Therefore, the majority of the NS interface area in Sample
4 has an additional tunneling barrier between the superconductor and
2DEG, namely the depletion region shown in
Fig.~\ref{cartoon1}(b). However, the type of NS junctions we form
by diffusion In into AlGaAs/GaAs are inhomogeneous enough that a
significant fraction of the sample can form a transmissive NS
interface of the type shown in Fig.~\ref{cartoon1}(a). The
conductance we observe in Fig.~\ref{anomloc} will be a parallel
combination of these two different types of NS junctions, as shown
schematically in Fig.~\ref{cartoon2}. For voltages away from $V=0$,
the slow variation of the background Giaever tunneling conductance
dominates the dI/dV curve. For voltages very close to zero bias, the
weak localization phenomena at the transmissive regions dominate and
leads to the observed conductance peak and dip.
\begin{figure}
\centps{dI-dVguess.eps}{30}
\setlength{\epsfysize}{7.0truecm}
\caption{Possible explanation for conductance dip around zero voltage
observed inside the zero bias conductance peak. Low transmission
regions of the interface give both the overall Giaever tunneling shape
of the dI/dV versus V and the zero bias conductance peak. A few high
transmission regions of the contact could produce the zero bias
conductance dip.}
\label{cartoon2}
\end{figure}
The peak at finite bias in Figs.~\ref{anomloc} was first observed by
Poirier et al.~\cite{poirier}, who called it the `finite bias
anomaly'. One problem with using the simulations of Mormorkos et
al.~\cite{} to explain a conductance dip around zero bias is that it
requires a high interface transmission, whereas the data of Poirier et
al.~\cite{poirier} (and our own data) show a Giaever tunneling
background (low average interface transmission). The inherent
inhomogeneity of supposedly planar superconducting In contacts to the
2DEG in AlGaAs/GaAs we have demonstrated in this paper overcomes this
difficulty. A few high transmission point emitters can produce the
conductance dip around zero bias, whereas the majority of the contact
can maintain low overall interface transmissivity. The weak localization
dip around zero bias can therefore peacefully coexist with a Giaever
tunneling background conductance.
The weak localization correction to the conductance of a ballistic NS
junction could have been more clearly observed in Sample 1, were it
actually present in that sample. Similarly Sample 2 (and several other
samples we measured) did not exhibit the weak localization correction
to the Giaever tunneling conductance. The exact impurity configuration
near a particular NS interface will determine whether or not the weak
localization correction to the conductance appears in any given
sample. Perhaps it is therefore not surprising that the weak
localization correction to the conductance can be observed only in a
fraction of the samples.
A different mechanism which splits the zero bias conductance peak in
NI$_1$NI$_2$S junctions was developed in Ref.~\cite{lesovik}. Weak
localization inside the middle N region produces the zero bias
conductance peak. If the two insulators I$_1$ and I$_2$ have two
different transmission coefficients, the zero bias conductance peak is
split as shown in Fig.~3 of Ref.~\cite{lesovik}. These two barriers,
having different transmissivity, is the same mechanism proposed by
Poirier et al.~\cite{poirier} to account for the `finite bias anomaly'.
Ref.~\cite{poirier} proposed a model which used the Schottky barrier at
the NS interface to produce I$_2$, and an impurity inside the
semiconductor as I$_1$. The spacing between I$_1$ and I$_2$ is L, a
random number set by the impurity configuration. The McMillan-Rowell
resonance nearest the Fermi level survives in the conductance of an
NI$_1$NI$_2$S junction upon averaging over different L, producing a
finite bias anomaly whose voltage is set by the average L.
The composite `point emitter' model for the contact developed in this
paper may also provide some support for this NI$_1$NI$_2$S model for
the `finite bias anomaly'. An electron moving through the 2DEG past a
point emitter would see that emitter as a scattering center,
equivalent to an insulating barrier. The distance between the emitters
in Sample 1 is of the order 100nm, less than the electron phase
coherence length. In Sample 1, therefore, could be regarded as a type
of NI$_1$NI$_2$NI$_3$~...~S junction. Each normal metal region N would
also be weakly localized. This model may also produce a finite bias
anomaly, but would require further numerical support. A two-dimensional
numerical simulation, where the electrons could actually move around the
point emitter scattering centers, would be required to confirm this
picture.
\section{Conclusions}
\indent
We have measured the differential conductance of superconductor-normal
metal junctions formed by diffusing Indium into AlGaAs/GaAs
heterostructures. In grows into a AlGaAs/GaAs heterostructure having a
[100] oriented surface preferentially along the \{111\}
crystallographic planes. Instead of a planar diffusion profile, we
therefore find that In forms `inverted pyramids' or point contacts to
the 2DEG. Supposedly `planar' superconducting In contacts to the
electron gas in an AlGaAs/GaAs heterojunction are therefore actually
composed of many point emitters. Correlating the contact
microstructure observed on different samples with the differential
conductance spectroscopy of the NS contact allowed us both to explain
many observed features in the conductance and to determine the
mechanism of superconducting (ohmic) contacts to the 2DEG in this
materials system.
For NS junctions annealed at a moderate temperature for a short times,
so that the In point contacts do not physically touch the 2DEG, we
obtain highly transmissive NS junctions. Due to the contact
inhomogeneity, the point emitters nucleate and grow at different rates
into the semiconductor. We observed wave interference between these
different superconducting emitters in transmissive NS junctions. For
identically prepared NS junctions annealed at higher temperatures and
for longer times, so that the In point contacts grow together and have
direct physical contact with the 2DEG, we obtain a lower transmission
NS interface and Giaever tunneling. This is due to a depletion layer
which forms around the In which directly touches the 2DEG. Further
annealing simply increases the effective strength of the interface
barrier between N and S, as regions of the In which previously were
not in direct physical contact with the 2DEG come in contact with the
2DEG.
Since the semiconductor forming N is disordered and has a reasonable
phase coherence length, weak localization corrections to the
differential conductance around zero bias voltage are also observed in
this materials system. This zero bias conductance peak is a correction
to Giaever tunneling which has been previously observed by several
other groups~\cite{kleins1,vanwees1}. We observed an additional dip
inside this zero bias conductance peak which develops in some samples
at low temperature~\cite{poirier}. One possible explanation for the
additional dip is due to contact inhomogeneity, where a small
percentage of the contact is a nearly ballistic NS interface while
most of the NS contact area remains in the tunneling limit. This
conductance dip around zero bias voltage is therefore possible
evidence for the predicted weak localization correction to the
conductance of ballistic normal metal - superconductor junctions in
Ref.~\cite{marmorkos1}. In any case, explanations for this `finite
bias anomaly' should account for the actual non-planar physical
structure of the superconducting contact.
\section{Acknowledgments}
\indent
We gratefully acknowledge support from the David and Lucile Packard
Foundation and from the MRSEC of the National Science Foundation under
grant No. No. DMR-9400415. We thank Tamer Rizk, Richard Riedel, Manoj
Samanta and Supriyo Datta for many useful discussions.
\vspace{0.1in}
$^1$ Present address: Intel Corporation, RN2-40, 2200 Mission College
Blvd. Santa Clara, CA 95052.
$^2$ Present address: Xilinx, 2100 Logic Dr., San Jose, CA 95124.
$^3$ Present Address: Yale University, Department of Electrical
Engineering, New Haven, CT 06520.
$^4$ Present address: Dept. of Physics, University of North Florida,
Jacksonville, FL 32224.
|
\section{Introduction}
The long awaited recent report \cite{KTev} on a clear observation of direct CP
violation in $K\to\pi\pi$ decays, ${\rm Re}(\epsilon'/\epsilon) = (28.0 \pm
3.0 \pm 2.6 \pm 1.0)\times 10^{-4}$, is the first evidence for the important
role played by penguin amplitudes in the phenomena of CP violation
\cite{Paschos}. $B$ decays are expected to provide a variety of CP asymmetry
measurements, as well as measurerments of certain combinations of rates, some
of which carry the promise of determining the angles of the unitarity triangle
\cite{review}, $\alpha, \beta$ and $\gamma$. This can test the commonly
accepted hypothesis that CP violation arises solely from phases in the
Cabibbo-Kobayashi-Maskawa matrix \cite{KM}. Let us review \cite{talk} a few of
the ideas involved in this study, paying particular attention to the
role of penguin amplitudes.
\begin{itemize}
\item {\bf$\beta$}:
In the experimentally feasible \cite{CDF} and theoretically pure example of
$B^0(t)\to J/\psi K_S$ the decay amplitude is real to a very high
precision. Theoretically \cite{SanBi}, the time-dependent mixing-induced CP
asymmetry measures the phase $\beta\equiv -{\rm Arg}V_{td}$ controlling
$B^0$-$\bar B^0$ mixing to an accuracy of 1$\%$ \cite{pen}.
\item {\bf$\alpha$}:~
$B^0(t)\to \pi^+\pi^-$ involves direct CP violation from the interference
between a dominant current-current amplitude carrying a weak phase $\gamma$
and a smaller penguin contribution, which ``pollutes" the measured $\sin\Delta
mt$ term in the time-dependent asymmetry \cite{pen}. A ratio of penguin to tree
amplitudes $|P/T|=0.3\pm 0.1$ in $B^0\to \pi^+\pi^-$ is inferred \cite{DGR}
from the measured rates \cite{CLEO} of $B\to K\pi$ dominated by a penguin
amplitude. Such a penguin contribution introduces a sizable uncertainty
\cite{MG} in the determination of $\alpha=\pi-\beta-\gamma$ in
$B^0\to\pi^+\pi^-$. Isospin
symmetry may be used \cite{GL} to remove this unknown correction to $\alpha$ by
measuring also the time-integrated rates of $B^{\pm}\to\pi^{\pm}\pi^0$ and
$B^0(\bar B^0)\to \pi^0\pi^0$. In the likely case that the decay rate into
$\pi^0\pi^0$ cannot be measured with sufficient precision, one can at least
use this measurement to set upper limits on the error in $\alpha$ \cite{GQ}.
Further out in the future, one may combine the time-dependence of $B^0(t)\to
\pi^+\pi^-$ with the U-spin related $B_s(t)\to K^+K^-$
to determine separately $\beta$ and $\gamma$ \cite{Duni}. This involves
uncertaities due to SU(3) breaking.
\item {\bf$\gamma$}:
The angle $\gamma$ is apparently the most difficult to measure.
It was suggested some time ago \cite{GLR} to obtain information about this
angle from charged $B$ decays to $K\pi$ final states by measuring the
relative phase between a dominant real penguin amplitude and a smaller
current-current amplitude carrying the phase $\gamma$. This is achieved by
relating the latter amplitude through flavor SU(3) \cite{GHLR} to the amplitude
of $B^+\to\pi^+\pi^0$, introducing SU(3) breaking in terms of $f_K/f_{\pi}$.
\end{itemize}
In the above two examples of determining $\alpha$ and $\gamma$, QCD penguin
amplitudes were taken into account in terms of their very general properties,
whereas electroweak penguin (EWP)
contributions were first neglected and later on analyzed in a model-dependent
manner \cite{ewp}. Such an approach relies on factorization and on
form factor assumptions \cite{model}, and involves theoretical uncertainties in
hadronic matrix elements similar to those plaguing
$\epsilon'/\epsilon$ \cite{Paschos}.
In the present report we will focus on recent developments in the
study of EWP contributions, which partially avoid these
uncertainties, thereby improving the potential accuracy of measuring
$\alpha$ and $\gamma$.
\section{Model-independent treatment of electroweak penguins}
The weak Hamiltonian governing $B$ decays is given by \cite{Buras}
\begin{equation}\label{H}
{\cal H} = \frac{G_F}{\sqrt2}
\sum_{q=d,s}\left(\sum_{q'=u,c} \lambda_{q'}^{(q)}
[c_1 Q_1 + c_2 Q_2]
- \lambda_t^{(q)}\sum_{i=3}^{10}c_i Q^{(q)}_i\right)~,
\end{equation}
where $Q_1=(\bar bq')_{V-A}(\bar q'q)_{V-A}, Q_2=(\bar bq)_{V-A}
(\bar q'q')_{V-A},~\lambda_{q'}^{(q)}=V_{q'b}^*V_{q'q},~q=d,s,~q'=u,c,t,
~\lambda_u^{(q)}+\lambda_c^{(q)}+\lambda_t^{(q)}=0$.
The dominant EWP operators $Q_9,~Q_{10}$ ($|c_{7,8}|\ll |c_{9,10}|$)
have a (V-A)(V-A) chiral structure, similar to the current-current operators
$Q_1, Q_2$. Thus, isospin alone
relates the matrix elements of these operators in $B^+\to \pi^+\pi^0$
\cite{GPY}
\begin{equation}\label{EWpi}
\sqrt2 P^{EW}(B^+\to \pi^+\pi^0)=\frac32 \kappa(T+C)~,~~~~~\kappa=\frac{c_9 +
c_{10}}{c_1 + c_2} = -0.0088~,
\end{equation}
where $T+C$ represents graphically \cite{GHLR} the current-current amplitudes
dominating $B^+\to \pi^+\pi^0$. Similarly, flavor
SU(3) implies \cite{GPY}
\begin{equation}\label{EW1}
P^{EW}(B^+\to K^0\pi^+) + \sqrt2 P^{EW}(B^+\to K^+\pi^0) =
\frac32\kappa (T+C)~,
\end{equation}
\begin{equation}\label{EW2}
P^{EW}(B^0\to K^+\pi^-) + P^{EW}(B^+\to K^0\pi^+) =
\frac32 \kappa (C-E)~.
\end{equation}
In the next three sections we describe briefly applications of these three
relations to the determination of
$\alpha$ and $\gamma$ from $B\to \pi\pi$ and $B\to K\pi$, respectively.
\section{Controlling EWP contributions in $B\to\pi\pi$}
The time-dependent rate of $B^0\to \pi^+\pi^-$ includes a term
$\sim\sin(2\alpha+
\theta)\sin(\Delta mt)$, where the correction $\theta$ is due to penguin
amplitudes \cite{GL}. Using isospin (\ref{EWpi}), the EWP
contribution to $\theta$, denoted by $\xi$, is found to be very small
\cite{GPY,BF}
\begin{equation}
\tan\xi = \frac{x\sin\alpha}{1+x\cos\alpha},~~
x\equiv \frac32\kappa
|\frac{\lambda_t^{(d)}}{\lambda_u^{(d)}}| = -0.013
|\frac{\lambda_t^{(d)}}{\lambda_u^{(d)}}|~,
\end{equation}
and is nicely incorporated into the analysis of Ref.~12 which determines
$\alpha$.
\section{$\gamma$ from $B^+\to K\pi$}
Using (\ref{EW1}), EWP terms are included in the triangle construction of
Ref.~15 \cite{NR2}
\begin{equation}
\sqrt2 A(B^+\to K^+\pi^0) + A(B^+\to K^0\pi^+) =
\tilde r_u A(B^+\to\pi^+\pi^0) \left(1 - \delta_{EW} e^{-i\gamma}\right)~,
\end{equation}
where $\tilde r_u = (f_K/f_{\pi})\tan\theta_c\simeq 0.28,~
\delta_{EW}=-(3/2)|\lambda^{(s)}_t/\lambda^{(s)}_u|\kappa \simeq 0.66\pm 0.15$.
This relation and its charge-conjugate permit a determination of $\gamma$
\cite{GLR,NR2}
under the {\it assumption} that a rescattering amplitude with phase
$\gamma$ can
be neglected in $B^+\to K^0\pi^+$. This amplitude is bounded
by the U-spin related rate of $B^{\pm}\to K^{\pm}\bar K^0$ \cite{Falk,GR,FL}.
Present limits are at the level of $20-30\%$ of the dominant penguin amplitude
\cite{GPY,Neubert}, and are expected to be improved to the level of 10$\%$.
In this case the rescattering effect, which depends strongly on
the final state phase difference $\phi$ between $I=3/2$ current-current and
penguin amplitudes,
introduces an uncertainty at a level of $15^{\circ}$ in the determination of
$\gamma$ if $\phi$ is near $90^{\circ}$ \cite{GP}. A considerably smaller
theoretical error \cite{Neubert} would be implied if this measurable phase is
found to be far
from $90^{\circ}$.
Other sources of errors in $\gamma$, such as SU(3) breaking, are
discussed elsewhere at this meeting \cite{Neubert,Flei}.
We note that in this determination of $\gamma$ SU(3) breaking does not occur
in the leading penguin amplitudes as it does in some other methods \cite{Duni}.
The phase $\gamma$ can also be constrained by measuring only charge-averaged
$B^{\pm}\to
K\pi$ rates. Defining
\begin{equation}\label{R*def}
R^{-1}_*=\frac{2[B(B^+\to K^+\pi^0) + B( B^-\to K^-\pi^0)]}
{B(B^+\to K^0\pi^+) + B(B^-\to \bar K^0\pi^-)}~,
\end{equation}
one finds using (\ref{EW1}) \cite{GPY,NR1}
\begin{equation}
R^{-1}_* = 1 - 2\epsilon \cos\phi (\cos\gamma - \delta_{EW}) +
{\cal O}(\epsilon^2, \epsilon^2_A, \epsilon\epsilon_A)~,
\end{equation}
where \cite{GLR,NR1} $\epsilon = \tilde r_u \sqrt 2
|A(B^{\pm}\to\pi^{\pm}\pi^0)/A(B^{\pm}\to
K^0\pi^{\pm})|\sim 0.24$, while $\epsilon_A$ is the suitably normalized
rescattering
amplitude. The resulting bound
\begin{equation}\label{const}
|\cos\gamma - \delta_{EW}| \ge \frac{|1-R^{-1}_*|}{2\epsilon}~,
\end{equation}
which neglects {\it second order} corrections, can be used to exclude an
interesting region around $\cos\gamma = \delta_{EW}$ if $R^{-1}_*\ne 1$ is
measured. Again, this would be very difficult if $\phi\simeq 90^{\circ}$. The
present value of the ratio of rates is \cite{CLEO} $R^{-1}_*=2.1\pm 1.1$.
\section{$\gamma$ from the ratio of $B^0\to K^{\pm}\pi^{\mp}$ to $B^{\pm}\to
K^0\pi^{\pm}$ rates}
Denoting this ratio of charged-averaged rates by $R$ ~\cite{FM},
one finds using
(\ref{EW2}) a constraint very similar to (\ref{const}) \cite{GPY,BF,FL}
\begin{equation}\label{const'}
|\cos\gamma - \delta'_{EW}| \ge \frac{|1-R|}{2\epsilon'}~
\end{equation}
where $\delta'_{EW}\sim 0.2\delta_{EW}\sim 0.13$ represents color-suppressed
EWP contributions, and \cite{GR}~ $\epsilon'\sim 0.2$ is the ratio of tree to
penguin amplitudes in $B^0\to K^+\pi^-$. In contrast to (\ref{const}), this
bound neglects {\it first order} rescattering effects, and the values of
$\delta'_{EW}$ and $\epsilon'$ are less solid than those of $\delta_{EW}$ and
$\epsilon$ in (\ref{const}). Eq.~(\ref{const'}) can exclude a region around
$\gamma=90^{\circ}$ if $R\ne 1$ is found. Presently \cite{CLEO} $R=1.07\pm
0.45$.
\section{Conclusion}
\begin{itemize}
\item In $B\to\pi\pi$ strong and electroweak penguins are controlled by isospin.
\item In $B\to K\pi$ strong penguins dominate and EWP are controlled by SU(3).
\item Interesting bounds on $\gamma$, in one case susceptible to rescattering
effects, are implied if the $B\to K\pi$ charge-averaged ratios of rates differ
from 1.
\item A precise determination of $\gamma$ from $B\to K\pi$ is challenging and
requires a combined effort involving further theoretical and experimental
studies.
\end{itemize}
\noindent
{\bf Acknowledgment}:
This work is supported by the United States $-$
Israel Binational Science Foundation under Research Grant Agreement 94-00253/3.
|
\section{Introduction}
Over the last several years we have learned a great deal about
supersymmetric gauge theories following the discovery
of dualities between string/M-theory and
supersymmetric gauge theories
\cite{bfs97,sus97,mald97}. Recently
this has been extended to conformal field theories without
supersymmetry \cite{klt99}.
Evidently, it would be desirable to have a deeper understanding
of supersymmetry breaking in order to bridge
the gap between the formulation of physics in a supersymmetric world, and
its more realistic counterpart, where no such symmetry is manifest.
One straightforward approach is to start with
a supersymmetric formulation, and then proceed to break
supersymmetry `softly' by adding appropriate mass terms.
The context within which we will consider this is a theory that has been
well studied
before: two dimensional SU($N$) gauge theory coupled to an adjoint Majorana
fermion
\cite{bdk93}. Interestingly, this theory is known to exhibit
supersymmetry at a particular value of the fermion mass, $m=m_{SUSY}$
\cite{kut93}.
This is believed to be a theory with two parameters $g$ and $m$,
both of which
have the
dimensions of mass. Since the only $g$ dependence is an overall $g^2$
factor in the
Hamiltonian the theory depends on one dimensionless parameter $X={m^2 \pi
\over g^2 N}$
and therefore all the bound state masses, in units of ${g^2 N \over \pi}$,
must be
determined in terms of the one parameter $X$. In this work, we provide
evidence that,
while this viewpoint is still correct,
there is still scope for an additional operator
(and associated coupling constant) that may be introduced
to improve convergence of
the DLCQ bound state masses towards their actual continuum values.
Of course, these continuum masses will be unaffected
by the presence of such an operator, but a judicious choice of
coupling will serve to improve the rate of convergence
of our numerical results.
Naively, when one adds a `soft' breaking term to the two
DLCQ formulations of the theory,
we appear to arrive at different spectra.
The two formulations we are alluding to are the
`Principle Value' (PV) and `Supersymmetric Discrete light Cone
Quantization' (SDLCQ), and
are discussed
in detail below. The question is whether these two spectra are truly
different or
simply rescalings of the same spectrum which become identical in the
continuum limit.
It would be very good news if they were in fact the same because
the PV prescription is
generally accepted as correct (\cite{pab85,bpp98}), while the SDLCQ approach
is known to converge more rapidly in general.
Actually, to understand the relation between these two schemes,
it is helpful to present a formulation that {\em interpolates} between the
PV and SDLCQ prescriptions by introducing an
additional operator \cite{alp99} and associated
coupling constant that we will call $Y$.
In particular, $Y=0$ will correspond to the PV prescription,
while $Y=1$ will imply the SDLCQ prescription. Intermediate
values for $Y$ will correspond to a `mixture' of the two schemes.
By diagonalizing the DLCQ Hamiltonian matrix, and extrapolating to
the continuum limit, we are able to solve for bound state
masses and wave functions at
different values of the fermion mass parameter $X$ and coupling
constant $Y$.
We shall show that the continuum bound state masses are independent
of the coupling $Y$, as expected from a scheme independent
prescription,
although the {\em rate of convergence}
towards the actual continuum mass will be significantly affected
by our choice for $Y$. In fact, it will turn out
that the value for $Y$ that arises naturally in the regularization
of supersymmetric theories (i.e. $Y=1$) provides the
best convergence towards actual continuum masses.
Thus, the supersymmetric formulation of DLCQ (SDLCQ),
corresponding to $Y=1$ -- first
highlighted in the work \cite{mss95} -- yields
a method for improving numerical convergence of DLCQ bound state masses
even for theories {\em without supersymmetry}.
To show this, we study the DLCQ bound state integral equations
at high resolution, which is made possible by
truncating the Fock space to two particles.
In particular, we show that the SDLCQ approach converges more uniformly
and rapidly for all values of $X$ that are
sufficiently far from the critical value $X=0$.
We remark that
at the supersymmetric point $X=1$, the SDLCQ prescription
preserves supersymmetry even in the discretized theory.
The advantages of such an approach have
been exploited in a study of a wide class of supersymmetric gauge theories
in two
\cite{alp98a,alp98b,alpp98,alpp99,alp99} and three dimensions
\cite{alp99b}.
\section{\bf Formulations of The Theories}
In this section we will consider the formulations of $1+1$ dimensional QCD
coupled to
adjoint Majorana fermions having arbitrary mass (see for example
\cite{bdk93}) in the light
cone gauge $A^+=0$. After eliminating
non-physical degrees of freedom by solving constraint equations,
the light--cone
components of total momentum are found to be:
\begin{eqnarray}
\label{momenta}
P^+&=&\int dx^- Tr(i\psi\partial_-\psi),\\
P^-&=&\int dx^-Tr\left(-\frac{im^2}{2}\psi\frac{1}{\partial_-}\psi-
\frac{g^2}{2}J^+\frac{1}{\partial_-^2}J^+\right) \nonumber \\
&=&\frac{m^2}{2}\int_0^\infty\frac{dk}{k}b^\dagger_{ij}(k)b_{ij}(k)+
\frac{g^2N}{\pi}\int_0^\infty\frac{dk}{k}\int_0^k dp \frac{k}{(p-k)^2}
b^\dagger_{ij}(k)b_{ij}(k) +\nonumber\\
&&\frac{g^2}{2\pi}\int_0^\infty
dk_1dk_2dk_3dk_4\left(\delta(k_1+k_2-k_3-k_4)
A(k)b^\dagger_{kj}(k_3)b^\dagger_{ji}(k_4)b_{kl}(k_1)b_{li}(k_2)+\right.\\
&&\left.\delta(k_1+k_2+k_3-k_4)B(k)(b^\dagger_{kj}(k_4)b_{kl}(k_1)b_{li}(k_2)b_{
ij}(k_4)-
b^\dagger_{kj}(k_1)b^\dagger_{jl}(k_2)b^\dagger_{li}(k_3)b_{ki}(k_4))
\right)\nonumber
\end{eqnarray}
with
\begin{eqnarray}
A(k)=\frac{1}{(k_4-k_2)^2}-\frac{1}{(k_1+k_2)^2},\nonumber\\
B(k)=\frac{1}{(k_3+k_2)^2}-\frac{1}{(k_1+k_2)^2}.
\end{eqnarray}
Here $x^\pm =(x^+ \pm x^-)/\sqrt2 $ and
$J^+_{ij}=2\psi_{ik}\psi_{kj}$ is the longitudinal component of the
fermion current. To avoid
introducing an additional mass scale in the theory we will write this in
term of mass operators:
$M^2=2P^+P^-$. It is well known that at the special value of fermionic mass
(namely
$m_{SUSY}^2=g^2N/\pi$) this system is supersymmetric \cite{kut93}. We will
use a dimensionless
mass parameter $X=\frac{\pi m^2}{g^2N}$, and the supersymmetric point is
$X=1$ and the masses
of all bound states will be quoted in units of
$g^2 N/\pi$.The supercharge is given by
\begin{eqnarray}
\label{sucharge}
Q^-&=&2^{1/4}\int dx^-tr(2\psi\psi\frac{1}{\partial_-}\psi). \nonumber \\
&=&\frac{i2^{-1/4}g}{\sqrt{\pi}}\int_0^\infty dk_1dk_2dk_3
\delta(k_1+k_2-k_3)
\left(\frac{1}{k_1}+\frac{1}{k_2}-\frac{1}{k_3}\right)\times\nonumber\\
&\times&\left(b^\dagger_{ik}(k_1)b^\dagger_{kj}(k_2)b_{ij}(k_3)+
b^\dagger_{ij}(k_3)b_{ik}(k_1)b_{kj}(k_2)\right),
\end{eqnarray}
\noindent Using the anticommutator at equal $x^+$:
\begin{equation}
\{\psi_{ij}(x^-),\psi_{kl}(y^-)\}=\frac{1}{2}\delta(x^- -y^-)
\end{equation}
it can be checked that at $m=m_{SUSY}$ the SUSY algebra
$\{Q^-,Q^-\}=2\sqrt{2}P^-$ is satisfied. In the DLCQ approximation the
system lives in a $x^-$ box
of length $L$ and one has to sums over discrete variables
$k^+ \ne 0$ instead of integrations in the above formulas. For periodic
boundary conditions (BC),
$k^+= n\pi/L $ where $n=1,2,\dots, K$ and
$K$ is called the resolution.
One formulation of DLCQ which we will denote as the
principal value (PV) prescription \cite{tho74}, treats the singularities of the
Hamiltonian using a PV prescription and can be formulated using
either anti- periodic or periodic BC. The anti-periodic boundary
condition must break the
supersymmetry at finite resolution because the fermions and bosons are in
different Fock
sectors. The PV prescription with periodic BC could in principle give
supersymmetric results
at finite resolution, although this is not the case.
In the PV prescription the
supersymmetry at $X=1$
is restored only in the decompactification limit ($K\rightarrow\infty$). This
restoration was
shown in \cite{bdk93} \footnote{They find that convergence is slower for
period BC}. The
Hamiltonian for this formulation will be referred to as
$P^-_{PV}$.
The prescription that preserves supersymmetry at finite
resolution will be called
SDLCQ. In SDLCQ one simply uses DLCQ to calculate the supercharges and then
uses
the super charges to calculate the Hamiltonian and longitudinal momentum
operator \cite{mss95}. Here we must use periodic BC because the supercharge
$Q^-$ is cubic in the fields, while the supercharge $Q^+$ is
quadratic.
The SUSY algebra is reproduced at a special value of
fermion mass and at
every finite
resolution the supercharge matrices give a representation of the super
algebra.
Both SDLCQ and $P^-_{PV}$ at $X=1$ give the same results as the resolution goes
to infinity \cite{alp98b} \footnote{However SDLCQ converges much faster}.
We now want to add
identical `soft' SUSY breaking terms (mass terms) to
these theories and study the resulting non-supersymmetric theory. Since
we already have a
mass term in $P^-_{PV}$ this only requires varying $X$, but for
SDLCQ this means explicitly adding a mass term.
It is very instructive to actually do the numerical calculation
differently and introduce a third formulation, $P_{SUSY}$ which includes
both SDLCQ
and $P^-_{PV}$ . We
have found the operator which is the difference between the
SDLCQ and the PV
formulation \cite{alp99}
\footnote{To date we have only found this operator for this particularly
simple theory but it
should be possible to find it for other theories as well.
The calculation of this operator involves a careful study of the
intermediate zero modes that
contribute to the square of the supercharge}. Thus if we add this operator to
the PV Hamiltonian it is now supersymmetric at every resolution and
produces exactly the same
mass and wave functions as SDLCQ. In the large $N$ approximation the
operator take the form.
\begin{equation}
\label{operator}
\frac{g^2NK}{\pi}\sum_n \frac{1}{n^2}
B^\dagger_{ij}(n)B_{ij}(n).
\end{equation}
Numerically, this operator does not alter the actual
continuum values observed in the PV approach
when $X=1$. In our
numerical formulation of
$P^-_{SUSY}$ we included this operator with an adjustable coupling $Y$. We
can now think of
$P^-_{SUSY}$ as a single theory in the coupling constant space
$(X,Y)$. The formulation we called PV corresponds
to setting $Y=0$ and allowing $X$ to be arbitrary,
while the prescription we call SDLCQ corresponds to
setting $Y=1$. In the following, we will present results for the lightest
bosonic bound states
as a function of $X$ and $Y$. For a few values of $X$ and $Y$ we will
truncate the Fock space
to allow only two
particles Fock states, which will permit
us to investigate the t'Hooft equation for higher resolutions than
would otherwise be possible.
\section{\bf `Soft' SUSY Breaking}
Our investigation of this theory indicates that at $X=1$ ( the
supersymmetric value of the fermion mass) the lightest
fermionic and bosonic bound states are degenerate with
continuum masses approximately $M^2= 26$ \cite{bdk93,alp98b}.
Using $P^-_{SUSY}$ we arrive at the same conclusion
for any value of $Y$.
Boorstein and Kutasov \cite{kub94} have investigated `soft' supersymmetry
breaking for small values of
this difference,
$X-1$ and they found that the degeneracy between the fermion and boson
bound state masses is broken
according to
\begin{equation} M^2_F(X) - M^2_B(X)= (1-X) M_B(1)+O((X-1)^3).
\label{linear}
\end{equation}
They calculated these masses using the PV prescription ($Y=0$) with
anti-periodic BC and found
very good agreement with the theoretical prediction. We have compared this
theoretical prediction
at $Y=1$ and we find
that eq (\ref{linear}) is very well satisfied. At resolution $K=5$, for
example, the slope is 4.76
and the predicted slope
$M_B(1)$ is 4.76. The indication is that this result is true for any value
of $Y$.
\begin{figure}[h]
\begin{center}
\epsfig{file=twoplots.eps, width=14cm}
\end{center}
\caption{(a) The contour plots of
$Y=Y(X)$ for the mass squared of the lowest bound state in units of $g^2 N/
\pi$ as a function of
$X=m \pi /g^2 N$ and Y (b)The contour plots of
$Y=Y(X)$ for the mass squared of the second lowest bound state in units of
$g^2 N/
\pi$ as a function of
$X=m \pi /g^2 N$ and Y (b) }
\end{figure}
In Fig. 1 we show the contour plots of the mass squared $M^2$
of the two lightest bosonic bound
states as a function of $X$
and $Y$ at resolution $K=10$. These contours
are lines of constant mass squared. Selecting a particular value of the
mass of the first bound
state then fixes a particular contour in Fig. 1a as a contour
of fixed mass, which we can write as
$Y=Y_p(X)$.
Interestingly, constructing the same contour plot
for the next to lightest bosonic bound state -- see
Fig. 1b -- yields contours that have approximately
the same functional dependence implied by Fig. 1a.
In fact, one obtains approximately the same contour
plots for the next twenty bound states (which is as far as we checked).
The simple conclusion is that the coupling $Y$ which
represents the strength of the additional operator
affects all bound state masses more or less equally.
This in turn suggests that at finite resolution, we can
smoothly interpolate between different values of fermion
mass $X$, and different prescriptions specified by the
coupling $Y$, without affecting too much the actual numerical spectrum.
Of course, in the decompactification limit $K \rightarrow \infty$,
such a dependence on $Y$ disappears, due to scheme independence.
Since the lightest bosonic bound state is primarily a
two particle state it is reasonable to truncate the
Fock basis to two
particle states. This will permit
very high resolutions, which will be needed
to carefully scrutinize any possible discrepancies between
the two versions of
'soft' symmetry breaking presented here.
In fact, we are able to study the theory
for $K$ up to 800. The
mass of the lowest state as a function of the resolution for various
values of $X$ and $Y$ are
shown in Fig. 2. Each converging pair
of lines -- which extrapolate the actual data points -- in Fig. 2
corresponds to different
values of fermion mass $X$. The top
upper curve in each pair runs through data points
that were calculated via SDLCQ (i.e. $Y=1$), while the
lower corresponds to the PV (i.e. $Y=0$) prescription
commonly adopted in the literature. We
find that each pair of
curves converge to the same point at infinite resolution,
although this may not be completely obvious
for the lowest pair in the
figure (corresponding to the critical mass $X=0$).
Away from
$X=0$, the SDLCQ formulation is fitted with a linear function of $1/K$,
while the PV formulation is fit
with a polynomial of
$1/K^{2\beta}$, where $\beta$ is the solution of $1-X/2=\pi \beta
Cot(\pi
\beta)$ \cite{van96}. It
now appears that SDLCQ not only provides more rapid convergence
for supersymmetric models, but
also for the
massive t'Hooft model, which is not supersymmetric.
For the massless case, the situation is reversed;
the SDLCQ formulation
converges slower. It is fit by a polynomial in $1/Log(K)$ and gives the
same mass at infinite
resolution as the PV formulation. This behavior
may be understood from the observation that
the wave function of this
state does not vanish at
$x=0$. We have looked closely at `small' masses,
such as $X=.1$, and one finds that both PV and SDLCQ
vary as a polynomial
in $1/K^{2\beta}$ at large resolution. Thus careful extrapolation
schemes must be adopted at small masses.
We therefore conclude that the continuum of
regularization schemes that interpolate smoothly between
the SDLCQ and PV prescriptions -- which we characterized
by the parameter $Y$ -- yield the same continuum
bound state masses, although the rate of
convergence of the DLCQ spectrum may be altered significantly.
This implies that
the contour plots observed in Fig. 1 eventually
approach lines parallel to the $Y$ axis, and the
sole dependence on the parameter $X$ is recovered.
\begin{figure}[h]
\begin{center}
\epsfig{file=convergenceplot.eps}
\end{center}
\caption{Mass of the of the lowest bound state in units of $g^2 N/ \pi$
calculated in the t'Hooft
model. The top pair is at $X=1$, the second is at $X=.5$,
and the bottom pair is at $X=0$ }
\end{figure}
\section{\bf Discussion}
The two dimensional gauge theory of adjoint Majorana fermions has been
studied extensively
\cite{bdk93,kut93,alp98b,kub94,anp98} and is known to be a theory with an
overall mass scale $g^2$,
and one real coupling -- the mass of the fermion -- which we
write as $X$ in our notation.
When one adds a `soft' supersymmetry breaking term, the supersymmetric
(SDLCQ) and principle
value (PV) prescriptions for regulating
the Coulomb singularity appear to give different bound state
masses at finite resolution.
We observed at finite resolution that these different
bound state masses may be smoothly connected -- in an approximate sense --
by introducing a new operator, and an associated coupling $Y$,
and then varying the couplings $X$ and $Y$ along an
appropriately chosen contour.
By truncating the Fock space to two particles,
we were able to study the DLCQ bound state equations up to
$K=800$, which we summarized in Fig. 2.
We concluded that after carefully extrapolating
the data, the different prescriptions yielded
identical continuum bound state masses.
Moreover, we observed that the SDLCQ prescription
improved convergence for sufficiently large values
of fermion mass.
Interestingly, since the
two-body equation studied here
for the adjoint fermion model
is simply the t'Hooft equation with
a rescaling of coupling constant,
we have arrived at an
alternative prescription
for regulating the Coulomb singularity
in the massive t'Hooft model
that improves the rate of convergence towards
the actual continuum mass.
Thus, a prescription that
arises naturally in the study of supersymmetric
theories is also applicable in the study of
a theory without supersymmetry.
We believe that this idea deserves to be exploited
further in a wider context of theories.
In particular, it is an open
question whether this procedure could provide a sensible approach to
regularizing softly broken gauge theories with bosonic degrees
of freedom, and in higher dimensions.
In any case, it appears that the special cancellations
afforded by supersymmetry -- especially in
the context of DLCQ bound state calculations --
might have scope beyond the domain
of supersymmetric field theory. This would be a crucial first step
towards a serious non-perturbative study of
theories with broken supersymmetry.
{\bf Acknowledgments}
\noindent The work was supported in part by a grant from
the US Department of Energy. The authors are grateful
for useful discussion with Brett Van de Sande.
|
\section{Introduction}
Foster and Michael \cite{FM} have studied numerically in lattice gauge models an interesting system consisting of a spinless static color octet source at the origin which modifies the gluonic field of pure color gauge theory in its vicinity. They call this system a ``gluelump'' and we follow their nomenclature. In \cite{FM} and several earlier publications referred to in \cite{FM} they have discussed the spectrum of this system. It is of interest to use this system as a testing ground for models of low energy QCD. Two models which have been much discussed in the literature are the bag model and the flux tube model and both may be applied to this system. We shall restrict ourselves here to the bag model. This would appear to be a particularly simple system to analyse in the bag model since the bag has a definite centre unlike the case of light quarks or gluons confined in a bag.
\\There are earlier discussions of the system in the bag model \cite{CS} \cite{Juge} with emphasis on the ground state configuration. We extend the earlier work to higher excited states and make a comparison with the lattice results. Physically the system would be of interest if heavy gluinos were to be produced experimentally and last long enough for them to have ther color neutralized. This was the motivation of the study by Chanowitz and Sharpe \cite{CS} who were interested in the ground state configuration of the system (which they called the ``glueballino''). Another physical connection is to the states of hybrid mesons in the limit of short distances between the quark and antiquark. Juge, Kuti and Morningstar \cite{Juge} have studied static hybrid potentials both on the lattice and in bag models and have shown that the bag gives a good representation of the lattice results at small quark-antiquark separation. The present discussion elaborates their work, as well as earlier work by Hasenfratz et al \cite{Has} and Ono \cite{Ono} at the special point $r=0$ and for very small values of $r$. We confirm in the bag model the results of Foster and Michael \cite{FM} for the quantum numbers of the low lying states, with plausible ordering $ J^{PC}=1^{+-}$,$1^{--}$, $2^{--}$,$3^{+-}$, $2^{+-}$.
\\The system we study has similarities to a glueball, which consists entirely of gluons. In the case of the {\it gluelump} one of the gluons is very heavy, and is located at the origin. The analogous heavy meson state, $Q{\bar{q}}$ has also been studied in the bag model \cite{Shuryak} and it was noted that the problem of centre of mass motion (and spurious excited states) disappears.
\section{Bag Model Calculation}
In the bag model, the system consists (in SU(3) of color) of an octet color charge fixed at the origin and one or several cavity ``gluons'' neutralizing the color charge in a spherical cavity of radius $R$. The one gluon states have lowest energy and we shall study them first, but will later estimate roughly the energies of the lowest two-gluon states. With a single gluon the energy of the system will have three contributions: a volume term $4\pi R^3 \Lambda/3$ where $\Lambda$ is the bag constant, the energy of the gluon mode, $k$, and $V^C$, the Coulomb interaction energy between the gluon and the octet charge at the origin. Initially we neglect $V_C$ so that the gluon in the bag is described by a free field.
\\The dynamics of a massless vector field confined to a cavity is well known \cite{Close} and the boundary conditions for the color electric ${\bf E}$ and magnetic ${\bf B}$ fields are ${\bf E \cdot r} = 0$ and ${\bf B \times r = 0}$ at $r = R$. The fields ${\bf E}$ and ${\bf B}$ are solutions of the (vector) wave equation which may be represented as ${\bf L}Y_{lm}j_l$ with $Y_{lm}$ spherical harmonics, $j_l$ spherical Bessel functions and ${\bf L}= {\bf r \times \nabla}$.
There are two sets of solutions of the vector wave equation, one in which ${\bf E} \propto {\bf L}Y_{lm}j_l$ and ${\bf B \propto \nabla \times} {\bf L}Y_{lm}j_l$, called TE modes (because ${\bf E \cdot r} = 0$ automatically) and the other set, the TM modes, in which ${\bf E}$ and ${\bf B}$ are interchanged. From these forms one may derive the equations which determine $kR$ through the boundary conditions. For the TE modes $kR$ is a solution of the equation ${\frac{d}{dr}(rj_l(kr)) =0}$ at $r=R$, while for the TM modes $kR$ is a solution of $j_l(kr) = 0$ at $r = R$. The lowest roots of these equations are as follows:\\
TE1: $\:kR= 2.744,\, 6.117,... $ for $J^P=1^+$\\
TE2: $\:kR= 3.870,\, 7.443,... $ for $J^P=2^-$\\
TE3: $\:kR= 4.973,\, 8.772, ... $ for $J^P=3^+$\\
TM1: $\:kR= 4.493,\, 7.72\,,... $ for $J^P=1-$\\
TM2: $\:kR = 5.763,\, 9.09\,,... $ for $J^P=2^+$\\
Since they are all single gluon modes they have $C=-1$.\\
Therefore the five lowest modes, in increasing order of $kR$, are
\[ J^P = 1^+ (kR=2.744),\: 2^- (kR =3.870),\: 1^- (kR=4.493),\: 3^+ (kR= 4.973),\: 2^+ (kR=5.763) \]
We note that the first three states ($1^+, 2^-, 1^-$) are the same as found on the lattice \cite{FM} but the order of the $2^-$ and $1^-$ states is reversed. However we have still to consider the color Coulomb interaction with the octet charge at the origin, and we shall see that this interaction may easily reverse the order of the $2^-$ and $1^-$ states, to bring agreement with the lattice results. We treat the color Coulomb interaction by perturbation theory. This is justified, for our parameter choice, by the small ratio of the energy shift to the spacing between levels of the same quantum numbers. (\cite{Shuryak})\\
The color Coulomb interaction between the central charge and the gluon field is a consequence of the non-Abelian nature of gluons. It is easy to compute it incorrectly! For example, if we were to take the volume integral of ${\bf E_1 \cdot E_2}$ where ${\bf E_1}$ is the field of the point color charge and ${\bf E_2}$ is
that of the gluon, we would get zero since since ${\bf E_1}$ is longitudinal while ${\bf E_2}$ is transverse, so that ${\bf E_1 \cdot E_2}$ vanishes. We first obtain the color charge density of the gluon field which is proportional to ${\bf E_2 \cdot E_2}$, average over the polar angles $\theta, \phi$ and integrate with $(Q/r)$ where $Q$ is the central charge. An earlier calculation of this interaction is reported by Ono \cite{Ono} who uses for the gluon the approximation of uniform color charge density inside the spherical cavity. In Ono's approximation all states shift by the same amount, so that we must go beyond this approximation to find the relative shift of levels.\\
To compute the (color) Coulomb interaction between the central octet charge and the gluon we use a ``confined'' Coulomb potential which obeys the boundary condition that the potential vanishes at the boundary of the bag, $V(R)=0$; so the potential of a central Abelian charge $q$ is $\frac{q}{4\pi}(r^{-1}-R^{-1})$. There is an additional factor of $-3$ associated with the color $SU(3)$. For consistency we also include the change in self-energy, due to the boundary conditions, of the central charge which also depends on $R$, and the self-energy of the gluon which is estimated in the approximation of averaging over angles. The constant term in the potential does not contribute to the final answer.
We shall illustrate the computation for the case of the lowest TE1 mode in which ${\bf E_{10}} = c{\bf L}Y_{10} j_1(kr)$ where $c$ is a normalization constant and we have chosen $(l,m) = (1,0)$ to simplify the calculation. We have immediately that $[{\bf E_{10}}]_z =0$ and, from angular momentum algebra,
\[ [{\bf E_{10}}]_x =-ic\frac{y}{r}\sqrt{\frac{3}{4\pi}}j_1(kr),\:\:
[{\bf E_{10}}]_y =+ic\frac{x}{r}\sqrt{\frac{3}{4\pi}}j_1(kr). \]
Therefore ${\bf E_{10}^2} = c^2(1-\cos^2 \theta)\frac{3}{4\pi} j_1^2(kr)$ and integrating over all angles one obtains $<{\bf E_1}^2>$ proportional simply to $j_1(kr)^2$, leading to the following expression for the color Coulomb interaction energy:
\[ V^C = -3 \alpha_s \frac{kR}{R} \frac{\int_0^{kR}x^2 (1/x - 1/(kR)) dx j_1(x)^2}{\int_0^{kR}x^2 dx j_1(x)^2} \]
where the factor $-3 = <{\bf F_1 \cdot F_2>_1}$ is the coupling of the color octets into a singlet and $\alpha_s$ is the color fine structure constant. The factor $kR$ in this case is 2.744, but the formula also applies to TE2, TE3 modes where we replace $j_1$ by $j_2$ or $j_3$ and use the apropriate value of $kR$. Incorporating also the self-energies detailed above, this leads to the following estimate for the first three TE modes:
\[ V^C(TE1) = -2.36 (-1.133) \alpha_s/R, \] \[ V^C(TE2) = -2.13 (-0.837) \alpha_s/R,\] \[V^C(TE3) = -2.01 (-0.680) \alpha_s/R \]
where the bracketed numbers correspond to neglecting the self energies. These coefficients were obtained by numerical integration of the spherical Bessel functions in the expression above. They diminish with the angular momentum $l$ as the color charge density shifts further from the origin with increasing $l$.\\
For the TM modes, where the color electric field is proportional to ${\bf \nabla \times} {\bf L }Y_{lm}j_l$, the computation is a little more tedious, and we find for the TM1 mode
\[ <{\bf E_1}^2> = c^2\left(\frac{3j_1^2}{r^2} + {{j_1^\prime}}^2 + \frac{2j_1\prime{j_1}}{r}\right) \]
and a similar expression with different coefficents for the TM2 mode, from which we compute, including self-energies (excluding self-energies in brackets):
\[ V_C(TM1) = -4.74 (-4.04)\frac{\alpha_s}{R},\:\: V_c(TM2) = -3.32(-2.40)\frac{\alpha_s}{R}. \]
The Coulomb interactions are larger than for the TE modes because the charge density in the case of the TM modes is closer to the origin.
\\We now have the ingredients to compute the energy of the bag by finding the extremal value of R. In this way one obtains
\[ E = \frac{4}{3} (4\pi \Lambda)^{\frac{1}{4}} (\alpha_{nj} - \kappa \alpha_s)^{\frac{3}{4}} = .79 (\alpha_{nj} - \kappa \alpha_{s})^{\frac{3}{4}} {\rm GeV} \]
where $\Lambda$ is the bag constant, $\alpha_{nj} =(kR)_{nj}$, $\kappa$ is the coefficent of $\alpha_s$ in the color Coulomb interaction. (For example, for the TE1 mode, $\alpha_{nj}=2.744$ and $\kappa =+2.36$). We have followed Juge et al \cite{Juge} and taken ${\Lambda}^\frac{1}{4} =.315 {\rm GeV}$, $\alpha_s = .23$, and so obtain:
\[ E(1+) = 1.43 {\rm GeV}, E(2-) = 1.97 {\rm GeV}, E(3+) = 2.44 {\rm GeV}\]
\[ E(1-) = 1.98 {\rm GeV}, E(2+) = 2.64 {\rm GeV},\]
We note that the states of $J^P = 2^-, 1^-$ are now essentially degenerate for this value of $\alpha_s$. For any larger values of $\alpha_s$ their order is reversed and the order of the first three states is now as found on the lattice \cite{FM}.\\
Before working out further one gluon states one should consider the position of two gluon states, which can also neutralize the central octet charge. The energy of a two gluon state has contributions from the energies of each of the gluons, their color electric interaction with the central charge, their mutual color electric interaction, and a color magnetic interaction between them. There are also the source and gluon self-energies as in the case of the one gluon states. The lowest such state is composed of two TE1 gluons, since TE1 is the lowest single gluon state. The only change which occurs relative to the case of the one gluon state comes from the fact that the interaction between two octet states (either gluon with central charge or between the two gluons, has a different color factor $<{\bf F_1 \cdot F_2}>_{\bf 8} = -3/2$, half the size of the factor $<{\bf F_1 \cdot F_2}>_{\bf 1} = -3$, since the two octets are combined in an octet rather than a singlet. The interaction of each gluon with the central charge is the same as in the single gluon case apart from this factor of $\frac{1}{2}$ but there are two gluons. The interaction energy between the two gluons is a little cumbersome to estimate, so we shall here make the approximation of assuming a uniform color charge distribution for each of the gluons for this evaluation. In this approximation the color Coulomb interaction of the two gluons with the central charge is just the same as that of the single gluon previously ($-1.13 \alpha_s/R$) and the interaction between the two gluons is
\[ \frac{1}{5}\frac{\alpha_s}{R}\left(-\frac{3}{2}\right)= -.3 \,\alpha_s/R
\]
Combining these results and the relevant self-energies, we estimate the two gluon states to have an energy of $.79 \,{\rm GeV}\: (5.5-2.4\alpha_s)^\frac{3}{4}$ $\simeq2.6 \,{\rm GeV}$. This estimate neglects the color magnetic interaction which will split the the various $J^P$ states from each other. With two $J^P=1+$ states one expects $0^+$, $1^+$ and $2^+$ states. Presumably the $0+$ state lies lowest with the $1+$ and $2+$ higher as with glueballs, though the glueball spectrum in the bag model does not correlate too well with that determined on the lattice. Therefore, one finds that the lowest states of the gluelump spectrum are $1^{+-}, 1^{--}, 2^{--}, 3^{+-}, 0^{+ +},2^{+-}$ with the relative positions of the $1^{--}$ and $2^{--}$ states and of the $0^{++}$ and $2^{+-}$ state not well determined by our estimates because of the uncertainty in $\alpha_s$.\\
\section{Hybrid Potentials At Small Quark-Antiquark Separation}
We next consider the effect of replacing the central octet charge by a pair of static color triplet and anti-triplet sources (in overall octet state) with the distance between the two sources being very small compared to the bag radius $R$. This system is equivalent to the addition of a small color octet dipole moment on top of the color octet charge. It is easy to see that the $J=1$ states will split into a ``molecular" $\Sigma$ state (coming form the $m=0$ component) and a molecular $\Pi$ doublet (coming from the $m=\pm1$ components) where $m$ is defined relative to a $z$-axis along the direction of the dipole. We argue below that the energy of these ``molecular" states varies quadratically as a function of the distance $\cal R$ between the triplet and anti-triplet with different coefficents for $\Sigma$ and $\Pi$ making the $\Pi$ doublet lower. It should be noted that we are neglecting at this point the color electrostatic interaction between the the triplet anti-triplet pair (which goes like $\alpha_s/{6\cal R}$) but this is the same in both $\Pi$ and $\Sigma$.
\\The action of the the dipole operator $\mu_z^{el}$ on the gluon wave function is equivalent to multiplication by the $z$-coordinate, where the $z$-axis is the axis of the dipole. This changes the multipolarity of the the gluon state, and in particular the parity of the gluon state flips. It is clear that all diagonal matrix elements vanish because of parity. Therefore the first non-zero contributions to the energy are in second order in the dipole operator, and therefore proportional to ${\cal R}^2$. We can evaluate approximately the coefficent of this quadratic term by using a single average energy denominator $\Delta E$, which we remove from the sum and use closure. In this way we obtain:
\[ E({\cal R}) = E(0) + \frac{9}{4} \alpha_s^2 <\frac{\cos^2 \theta}{r^4}>\frac{1}{\Delta E} {\cal R}^2 \]
where the expectation value is over the lowest state TE1,$m$ in the case of the ground state. One obtains two different values for $m=0$ and $m=\pm 1$. For the angular integrals we obtain $(2/5)$ for the $\Pi$ states ($m=\pm1$) and $1/5$ for the $\Sigma$ state ($m=0$). Recalling that for the ground state the energy denominator $\Delta E$ is negative, we find that the lower molecular state is the $\Pi$ state and the next excitation is the $\Sigma$, in agreement with the lattice computations of \cite{Juge}. Evaluating numerically the integrals over spherical Bessel functions we estimate the splitting between the $\Sigma$ and $\Pi$ states to be about $0.2\left(\frac{\cal R}{R_{\rm bag}}\right)^2$ in GeV, which is similar to the lattice estimate at short distances. \cite{Morning}
\section{Concluding Remarks}
Our bag model results for the order of levels in the gluelump spectrum are gratifyingly close to those of the lattice. The actual magnitude of the splitting between the lowest two levels is given as $.35$ GeV in \cite{Michael2}. With the bag parameters we have chosen it is somewhat larger, $.54$ GeV. It is also amusing that our crude bag model estimate of the $\Sigma-\Pi$ splitting of the lowest hybrid potentials also agree qualitatively with the lattice results. \cite{Michael3}\\
Some lattice gauge results for the gluelump spectrum also exist in the case of $SU(2)$ of color \cite{Michael3}. The corresponding bag model calculation differs from the above only by the change in group theory factors. Normalizing to the string tension which in the bag model is given by
\[ SU(3): \sigma = \sqrt{\frac{32 \pi \alpha_s \Lambda}{3}} \],
\[ SU(2): \sigma = \sqrt{6 \pi \alpha_s \Lambda} \]
and noting that the ``color Coulomb'' potential has about the same strength in SU(2) as in SU(3), we have \[ \Lambda(SU(2))=\Lambda(SU(3)),\:\:\frac{3}{4}\alpha(SU(2))\simeq\frac{4}{3}\alpha(SU(3)) \]
so that there is no appreciable change in the bag model spectrum in going from $SU(3)$ to $SU(2)$ (apart from the fact that the two gluon state must now have $C=-$ only). There is a suggestion from the lattice \cite{Michael2} that the separation in $SU(2)$ may be less than in $SU(3)$.\\
\acknowledgements{We thank PPARC for financial support and GK also thanks NSERC, Canada for support.}
|
\section{Introduction}
One of the main challenges of thermal QCD is to get reliable numbers.
Though the gauge coupling may be small, Linde's argument\cite{linde} tells us
that perturbation theory will fail. The powerlike infrared divergencies
one meets in perturbation theory will off-set the powers of the
coupling constant. At what order in perturbation theory this will
happen
depends on the observable in question. For the free energy this
happens when the static sector starts to dominate, and a
simple dimensional argument shows this will happen at $O(g^6)$. For
the
Debye mass Linde's phenomenon starts already at next to leading order.
So the problem is certainly not academic! One should bear in mind that
Linde's argument does not deny the existence of a perturbation series.
It says that from a certain order on the coefficients are no longer
obtained by evaluating a finite number of diagrams of a given loop order.
So we are faced with evaluating non-perturbative effects from the
three dimensional sector defined by the static configurations.
It was realized some time ago~\cite{dimred} that one could take the
static part of the 4D action combined with induced effects by the
non-static configurations. This theory gives at large distances the
same physics as
the 4D theory, and has the advantage of relatively straightforward
lattice simulations\cite{debye}. In section 2 we discuss the relation between
the 4D and the 3D theory. In particular we show how the phase diagram of the 3D
theory has a remarkable property: the curve of 4D physics, and the critical
curve as determined by perturbation theory do coincide to one and two loop order.
However, perturbation theory has no reason to be trustworthy in determining
the critical curve, and this is probably the reason why the fit to the numerical
determination is problematic.
In section 3 we discuss the physics of the domain wall in some detail.
\section{Effective 3D action and symmetry in 4D}\label{sec:eff3dact}
Construction of the effective action proceeds along familiar lines.
In the case of QCD with $n_f$ quarks its form is given by integrating out
the heavy modes of $O(T)$:
\begin{equation}
S_{3D}=S_{YM,n=0}+S_{ind}
\label{eq:s3d}
\end{equation}
The first term is the static sector of the pure Yang-Mills theory in 4D
with coupling constant $g_3=g\sqrt T$.
The second term in eq.~\ref{eq:s3d} must contain the symmetries of the
original QCD action, as long as they are respected by the reduction
process.
So we expect the induced action to be of the form:
\begin{equation}
S_{ind}=V(A_0)+ \mbox{ terms involving derivatives}
\end{equation}
$V(A_0)$ should be invariant under static gauge transformations,
C, CP ($A_0\to -A_0^T$) and this reduces it to a sum of traces of even
powers of $A_0$:
\begin{equation}
V(A_0)=m^2TrA_0^2+\lambda_1(TrA_0^2)^2+\lambda_2 TrA_0^4+....
\label{eq:va0}
\end{equation}
Only one independent quartic coupling survives for SU(2) and SU(3).
We take it to be ${TrA_0^2}^2$.
Note that we lost a symmetry present in the 4D action for gluons
alone, and less and less conserved when quarks get lighter and
lighter: Z(N) symmetry.
Remember from the lattice formulation of pure Yang-Mills that one can multiply at a given time slice in the original 4D action all links in
the time direction with a factor $\exp{\pm i\ {2\pi\over 3}}$. This will not
change the form of the action, but will change by the same factor the value of the Wilson
line $P$ wrapping around the periodic time direction:
\begin{equation}
P(A_0)={\cal {P}}{\exp{i\int A_0 d\tau}}
\end{equation}
Clearly in eq.~\ref{eq:va0} this symmetry has gone. Apparently the
reduction process does not respect
$Z(3)$ symmetry! The reason for this is twofold:
i)the reduction process
does not include the static modes.
ii)the values of $A_0/T$ in the effective action are order $g$, whereas
the Z(N) symmetry equates the free energy in $A_0 /T$ and $A_0/T+O(2\pi /3)$.
To understand this better -- and to prepare the way for the discussion
of the domain wall observable in the last section \ref{sec:domain} -- we
recall some familiar facts in 4D for SU(3).
\subsection{Z(3) symmetry and domain walls in 4D gauge theory}{\label{sec:domain1}}
The free
energy $U$ as a function of the Wilson line invariants $TrP,TrP^2$ is naturally
defined
through:
\begin{equation}
\exp{-{VU(t_1,t_2)\over T}}=\int DA_0 D\vec A \ \delta(t_1-\overline{ TrP})
\delta(t_2-\overline{TrP^2}) \ \exp{-{S(A)\over{g^2}}}
\label{eq:wilsonline}
\end{equation}
where $\overline{TrP}$ is the normalized space average of the trace over the volume $V$.
A natural parametrization of the parameters $t_1$ and $t_2$ suggests itself: define the
phase matrix $\exp{iC}$ with $C$ being a traceless diagonal 3x3
matrix with entries $C_i \ ,(i=1,2,3)$ and $\sum_i C_i=0$, because we have
SU(N), not U(N).
Consider pure Yang-Mills. A gauge transformation that is periodic
modulo
a phase in Z(3) will {\it only} change the arguments in the delta
functions in eq.~\ref{eq:wilsonline}.
Hence the potential $U$ has degenerate minima in all points
of the
C-plane, where $\exp{iC}=1$, or $\exp{\pm i2\pi/3}$. This is called Z(3)
symmetry (and the degeneracy is lifted by the presence of quarks).
This statement is independent of perturbation theory. In fact the
potential
in eq.~\ref{eq:wilsonline} has been computed in perturbation theory
including two loop order. And this potential includes the static modes.
Propagators acquire a mass proportional to the phases $C$, because it acts like a VEV of the adjoint Higgs $A_0$.
Hence, for small $C$, eventually Linde's argument will apply and
the perturbative evaluation becomes impossible.
For SU(3) the direction in which the Wilson line phase causes minimal
breaking is in the hypercharge direction $C={1\over
3}diag(q,q,-2q)$. Minimal breaking means the maximal number of
unbroken massless excitations, that do not contribute to the
potential. Hence this is at the same time the valley through which
the system tunnels from one minimum to the next. In this "q- valley"
the combined 1 and 2 loop result is exceedingly simple:
\begin{equation}
U^{(1)}+U^{(2)}={4\pi^2\over 3}T^4(N-1)\left(1-5
{g^2N\over{(4\pi)^2}}\right)q ^2(1-q)^2
\label{eq:pot4d2q}
\end{equation}
For use in the reduced theory we isolate the static part of the one and two loop
contribution in the q-valley from eq.~\ref{eq:pot4d2q}:
\begin{equation}
\left( U^{(1)}+U^{(2)}\right)_{(n=0)}
=-T^4(N-1){4\pi^2\over 3}\left(2q^3 +3{g^2N\over{(4\pi)^2}}q^2\right)
\label{eq:un0}
\end{equation}
Note that the two loop contribution is quadratic in q in contrast to
the one loop which is cubic. The two-loop cubic part in eq.~\ref{eq:pot4d2q} comes from a combination of static and non-static modes.
If we prepare the 4D system conveniently this symmetry will give rise to
domain walls. Profile and energy of these wall have been computed
semi-classically a long time
ago\cite{bhatta}. The
method
of twisted boundary conditions triggers walls and is most economic computerwise. We will discuss them
in the context of the lattice formulation in section~\ref{sec:domain}.
Be it enough to mention that these boundary conditions force the
Wilson lines
to change by a Z(N) phase in going from one side to another side of
the volume in some a priori fixed space direction. This will trigger a
wall profile for the loop in this direction.
It is the long range behaviour of this profile that contains the
information on the Debye mass. To one loop order this behaviour comes
entirely from the slope of the potential, see above. But to two loop
order we have to take the one-loop renormalization of the gradient part of the
Wilson line phase into account, and this suffers the Linde effect: there is an
infinity
of many-loop diagrams contributing to the gradient part. So to next to
leading order there
are already non perturbative effects in the long range tail of the wall, and hence in the Debye mass, as we mentioned earlier.
On the other hand we know that the
effective 3D action correctly reproduces the large
distance behaviour of the 4D theory. So a 3D projection of the twist
should produce a wall with the same tail as the 4D one. The inside of
the wall in both formulations may be quite different but the inside
is anyway computable by perturbation theory.
\subsection{3D action and 4D physics}{\label{sec:4dphysics}}
The parameters of the 3D theory ($m^2$ and $\lambda\equiv\lambda_1+\lambda_2$
for SU(3))
in eq.~\ref{eq:va0} can be calculated in perturbation
theory by integrating out all modes in a path integral except the
mode $A_\mu (\vec x, n=0)$. To one loop order we have the well known
result for the Debye mass and for the four point coupling $\lambda$.
All higher order terms have a coefficient zero~\cite{these}. To two
loop order one has to take care not only of the two loop graphs, but
also of the 1-loop renormalization of the
three dimensional gauge coupling $g_3$ and the renormalization of the
$A_0$ field in the gradient terms. The latter renormalization is taking care of gauge
dependence in the two loop graphs.
The result~\cite{loop} in the $\overline{MS}$ scheme is that both parameters are
expressed in the renormalized 4D coupling $g(\mu )$ where $\mu $ is the
subtraction point. Eliminating the 4D coupling gives for the
dimensionless quantities $x\equiv {\lambda\over{g_3^2}}$ and
$y={{m^2} \over{g_3^4}}$ the result for N=3:
\begin{equation}
xy_{4D}={3\over{8\pi^2}}(1+{3\over 2}x)
\label{eq:physline}
\end{equation}
whereas for N=2:
\begin{equation}
xy_{4D}={2\over{9\pi^2}}(1+{9\over 8}x)
\label{eq:physlinebis}
\end{equation}
Note the absence of explicit $\mu$ dependence in this relation. The
variable $x$ has a $\mu\over T$ dependence such that as T becomes large
$x$ becomes small.
In conclusion, it is along this line that we have to simulate the 3D
system,
in order to get information about the 4D theory. Before we do this, we
still have to settle an important question: where are -- in the $xy$
versus $x$ diagram -- possible phase transitions?
\subsection{Phase diagram of the 3D theory}{\label{secphase}}
To get the phase diagram we must first decide what order parameters to
take. In the case of SU(3) there are two: $TrA_0^2$ and $TrA_0^3$. Strictly speaking, only the latter is an order parameter, since it flips sign under C.
We will study the analogue of eq.\ref{eq:wilsonline}:
\begin{equation}
\label{effaction}
\exp{-VS_{eff}(D,E)}=\int
DA\delta\left(g_3^2D-\overline{TrA_0^2}\right)
\delta\left(g_3^3E-\overline{TrA_0^3}\right)\exp{-S}
\label{eq:effactiondef}
\end{equation}
Again as for the Wilson line we parametrize D and E in terms of $D=Tr\left[ C^2 \right]$
and $E=Tr\left[ C^3 \right]$ respectively.
Let us first state the result one gets for $S_{eff}$ to one and two
loop order:
\begin{equation}
S_{eff}={U(n=0)\over T}\hskip1cm \hbox{one and two loop only}
\label{eq:static}
\end{equation}
The one and two loop result equals the static part of the
4D Z(3) potential,eq.~\ref{eq:wilsonline}! This static part was
explicitely written in the q-valley, eq.\ref{eq:un0}. It has to be
added to the tree result and one gets in terms of the dimensionless
variables x and y for N= 2 or 3 colours, absorbing a factor $2\pi$ in q:
\begin{equation}
{S_{eff}\over{g_3^6}}= y\left({N-1\over N}\right)q^2
+x\left({N-1\over N}\right)^2q^4
-(N-1)\left({1\over{3\pi}}q^3 +{N\over{(4\pi)^2}}q^2\right)
\label{eq:effaction}
\end{equation}
The question is now: for what values of x and y we have degenerate
minima for q? Keeping only the 1 loop result cubic in q we see that
it must be of the
order of magnitude of the quartic term of the tree result to get a
second degenerate minimum. So q must be of
$O({1\over x})$ in that minimum. Thus the quadratic two loop result
contributes $O(x)$ less.
From eq.\ref{eq:effaction} we find the potential develops two
degenerate minima for N=3 when:
\begin{equation}
xy_{c}={3\over{8\pi^2}}(1+{3\over 2}x)
\label{eq:crit}
\end{equation}
For N=2:
\begin{equation}
xy_{c}={2\over{9\pi^2}}(1+{9\over 8}x)
\end{equation}
This is important: slope and intercept of the physics
line~\ref{eq:physline} are identical with those of the critical
line~\ref{eq:crit}, at least if we can take the low order loop results
for the critical line seriously. This was numerically found in ref.\cite{polonyi,loop}
The intercept equality is just due to the Z(N) potential in 4D and
the effective potential $S_{eff}$ in 3D being {\it identical} to one loop.
But to two loop order this simple explanation is no longer true. The
cubic term in eq.\ref{eq:pot4d2q} is appearing also in the two loop
result, but not in the two loop result for the 3D effective action. It
is however true that also in 2 loops the leading contribution is the
static part of the Z(N) potential, eq.\ref{eq:un0}.
\subsection{Saddle point of the effective potential in 3D}\label{sec:saddle}
In this subsection we will investigate in more detail the
computation
of the 3D effective potential. The saddle point is found by admitting
$A_0$ fluctuates around a diagonal and constant background B:
\begin{equation}
A_0=B+Q_0
\end{equation}
whereas the spatial gauge fields fluctuate around zero:
\begin{equation}
A_i=Q_i.
\end{equation}
One then goes through the usual procedure of expanding the effective
action
\ref{eq:effactiondef}. The equations of motion fix the background B to
be equal to the matrix C, and the part quadratic in the fluctuations
will not contain any reference to the Higgs potential $V(A_0)$.
This is clear because the quadratic constraint tells the mass
term not to fluctuate. Only the Higgs component parallel to C ,
$TrCQ_0$, has a mass term due to the Higgs potential, $4\lambda
TrC^2$. So apart from this the quadratic part comes entirely from
the static part of the 4D action.
We can make a convenient gauge choice, namely the static form
of the covariant gauge fixing:
\begin{equation}
S_{gf}=Tr\left([ig_3 B,Q_0]+\partial_kQ_k\right)^2
\label{eq:fixing}
\end{equation}
This gives propagators which are precisely the static version of the
propagators appearing in the Wilson line potential~\ref{eq:wilsonline}.
Only the component $Q_0$ parallel to C is the exception: its propagator
has a mass from the Higgs potential and can be written as the the sum of the
static propagator and a remaining part (``massive'')
containing the mass term:
\begin{equation}
{1\over {\vec p}^2}+ {1\over {\vec p}^2+4\lambda TrC^2}-{1\over {\vec p}^2}
\label{eq:prop}
\end{equation}
The static propagator dominates in diagrams over the rest.
The massive propagator will give rise to half integer powers
of x in the perturbative expansion of the potential; gauge couplings contribute $O(1)$ in dimensionless units, whereas Higgs couplings contribute
$O(x)$.
As long as we are interested in intercept and slope of the critical curve,
it follows that only the static part of the Feynman rules contributes.
Hence the result \ref{eq:static}.
Let's from now on work in the q-valley where we evaluate the effective
action \ref{eq:effaction}.
Then two remarks are crucial:
i)The broken minimum occurs for $q=O(1/x)$. Power counting then reveals
that from $O(x^{3/2})$ on an infinite number of diagrams contributes to
each order.
ii)From five loop order on, the potential starts to develop poles in
$q=0$.
We are bringing this up, because insisting on the low order
result \ref{eq:crit}
and fitting numerically the coefficients of $x^{3/2}$ and higher order gives an
unexpected result: the numerical coefficients are orders of magnitude larger\cite{loop} than
the first two in \ref{eq:crit}. In fig. 1, taken from ref.~\cite{loop},
the situation is shown. Only for very small x the critical and the 4D physics line
are allowed to become tangent. It seems that this constraint
affects the quality of the fit.
Dropping it altogether necessitates numerical determination of
transition points at $x\le 0.04$.
\begin{figure}[h]
\begin{center}
\hskip 0.truecm
\epsfbox{fig.eps}
\caption{Phase diagram in the SU(2) case, from ref. 9. The straight line
is the 4d $\rightarrow$ 3d curve of eq.\ref{eq:physlinebis}. The thick line is a 4th order
fit to the data.
The dashed line marks the region where the transition turns into a cross-over.}
\end{center}
\end{figure}
\section{Debye mass from a 3D domain wall}{\label{sec:domain}}
After this long discussion of where the physics line lies with respect
to the critical curve we have to come to grips with the domain wall
method.
The idea here is extremely simple and has been explained elsewhere\cite{altesbron}.
Twisted boundary conditions in 4D~\cite{groe} have a very simple and
intuitive form in the reduced theory.
Remember that a twisted plaquette in the time-space direction is of the
form $Tr(1-\Omega U(P))$, with $\Omega=\exp{i2\pi/N}$.
Thus intuitively one would say that all one has to do in the reduced
action is to modify the kinetic part of the Higgs field by the twist,
because that's what the plaquette in the time space direction is
reducing to.
In the next subsection we work out this idea in more detail.
\subsection{Construction of the wall}
In this section we want to make more precise the
action that defines the wall.
We follow the notation of ref.\cite{boucaud}, specifically that of
hep-lat/9811004 and write the kinetic part of the action as:
\begin{eqnarray}
{\cal L}_{kin} & = &
{36 \over \beta} \sum_{\vec x} Tr \left[ A^2({\vec x}) \right]
\ - \
{12 \over \beta} \sum_{{\vec x},j}
Tr \left[ A({\vec x}) U_j({\vec x}) A({\vec x}+ a{\vec e_j} ) U_j^+({\vec x})
\right]\nonumber
\\
& = &
{12 \over \beta} \sum_{{\vec x},j}
Tr \left[
{1\over 2} \left( A^2({\vec x})+ A^2({\vec x}+ a{\vec e_j})\right)
- A({\vec x}) U_j({\vec x}) A({\vec x}+ a{\vec e_j} ) U_j^+({\vec x})
\right]
\nonumber\\
\end{eqnarray}
where $\vec x$ is a vector with three components $(x,y,z)$.
\noindent
Consider the following expression:
\begin{eqnarray}
{\cal X } \ = \
1 \ - \ {1\over {\rm N} } {\cal R}e \ Tr \left[
e^{{\rm i} \alpha A(x,y,0)} U e^{-{\rm i} \alpha A(x,y,1)} U^+ \right]\nonumber
\label{eq:X}
\end{eqnarray}
If $\alpha A$ is small we get:
\begin{equation}
{\cal X }=
{1\over {\rm N} } \alpha^2 Tr \left[
{1\over 2} A^2(0) + {1\over 2} A^2(1 ) - A(0) U A(1) U^+\right]\nonumber
\end{equation}
This is precisely the kind of expression that appears in eq. (19).
>From this follows the expression for the modified kinetic energy in
the plane $(x,y,0)$:
\begin{eqnarray}
{\cal L}_{kin}^{mod} =
{12 \over \beta}
\kern -.4 truecm \sum_{{\vec x},j \atop (z,j) \ne (0,3) }
\kern -.3 truecm
Tr \left[
{1\over 2} \left( A^2({\vec x})+ A^2({\vec x}+ a{\vec e_j})\right)
- A({\vec x}) U_j({\vec x}) A({\vec x}+ a{\vec e_j} ) U_j^+({\vec x})
\right]&&
\nonumber\\
+ \ {12 \over \beta}\ { {\rm N} \over \alpha^2}\
\sum_{x,y}
\left\{
1 - {1\over {\rm N} } {\cal R}e \ Tr \left[
e^{{\rm i} \alpha A(x,y,0)}
U_3(x,y,0) e^{-{\rm i} \alpha A(x,y,1)} U_3^+(x,y,0) \right]
\right\}&&\nonumber\\
\end{eqnarray}
\noindent
So all we need is to put a twist $\Omega \ \in Z({\rm N})\ $ in order
to get a wall:
\begin{eqnarray}
{\cal L}_{kin}^{wall} =
{12 \over \beta}
\kern -.4 truecm \sum_{{\vec x},j \atop (z,j) \ne (0,3) }
\kern -.31 truecm
Tr \left[
{1\over 2} \left( A^2({\vec x})+ A^2({\vec x}+ a{\vec e_j})\right)
- A({\vec x}) U_j({\vec x}) A({\vec x}+ a{\vec e_j} ) U_j^+({\vec x})
\right]&&
\nonumber\\
+ {12 N \over \beta \alpha^2}
\sum_{x,y}
\left\{
1 - {1\over {\rm N} } {\cal R}e \left( \Omega \ Tr \left[
e^{{\rm i} \alpha A(x,y,0)}
U_3(x,y,0) e^{-{\rm i} \alpha A(x,y,1)} U_3^+(x,y,0) \right]
\right) \right\}&&\nonumber\\
\end{eqnarray}
\noindent
What is now the actual value of $\alpha$ to use?
We recover the kinetic term in the continuum if we relate the field
$A$ on the
lattice to the field $A_{cont}$ in the continuum by the relation:
\begin{eqnarray}
A \ = \ { A_{cont} \over g_3}\nonumber
\end{eqnarray}
\noindent
This is not the usual normalization for the lattice fields.
Usually we have: $A_{latt} \ = \ a g_3 A_{cont}$ and
$A_{latt} \ \rightarrow \ 0 $ in the continuum limit.
Here this is not anymore the case.
Remember that to expand the modified action we had to suppose that
$\alpha A$
was small. To enforce this condition it seems natural to put~: $ \alpha
\ = \ a g_3^2$; in this manner terms of the kind $e^{{\rm i} \alpha A} $
become $e^{{\rm i} a g_3^2 A} $. That is to say, they become of the
usual sort~: $e^{{\rm i} a g_3 A_{cont}} $.
With this choice the term in the exponential indeed goes to zero as
the lattice spacing goes to zero, so:
\begin{eqnarray}
\alpha \ \equiv \ a g_3^2 \ = \ { 6 \over \beta}
\nonumber
\end{eqnarray}
In the end we obtain as final expression for the kinetic part of the
action supporting the wall:
\begin{eqnarray}
{\cal L}_{cin}^{wall}
=
\displaystyle {12 \over \beta}
\kern -.4 truecm \sum_{{\vec x},j \atop (z,j) \ne (0,3) }
\kern -.31 truecm
Tr \left[
{1\over 2} \left( A^2({\vec x})+ A^2({\vec x}+ a{\vec e_j})\right)
- A({\vec x}) U_j({\vec x}) A({\vec x}+ a{\vec e_j} ) U_j^+({\vec x})
\right] &&\nonumber
\\ \nonumber
& &\\ \nonumber
\displaystyle + \beta \sum_{x,y}
\left\{
1 - {1\over {\rm N} } {\cal R}e \left( \Omega \ Tr \left[
e^{{\rm i} {6 \over \beta} A(x,y,0)} U_3(x,y,0) e^{-{\rm i}
{6 \over \beta}A(x,y,1)} U_3^+(x,y,0) \right]
\right) \right\}&&\\ \nonumber
& & \\
\label{eq:twistkin}
\end{eqnarray}
\subsection{Excitations of the wall}{\label{subsec:excwall}}
Now the system with the wall is defined by adding the 3D gauge field action and
the Higgs potential V(A) to eq.~\ref{eq:twistkin}. Let us call the
resulting twisted action $S_t$.
Both twisted and untwisted action have periodic boundary conditions.
When we compute the average of an observable $O$ in the twisted box
we average the observable over the $(x,y)$ plane at the point $z$,
written as $\overline{O(z)}$, and compute in the twisted box (action $S_t$).
It is quite trivial to relate this average to the correlation of the wall
and $O$ in the untwisted box (action $S$):
\begin{equation}
\langle \overline{O(z)}\rangle_{S_t}=\langle\exp{-(S_t-S)} \overline{O(z)}\rangle_{S}
\label{eq:twistav}
\end{equation}
There is no difference between the two actions except at $z=0$, at the location of
the wall.
The twist is C and P odd, but T even.
This means we can expect a signal for the Debye mass by taking
any observable $O$ C odd (a necessary condition\cite{arnold}). Whatever operator gives the lowest mass in the
correlation \ref{eq:twistav} is the preferred one. Thus one and
the same updating with the twisted box can be used for various operators.
\section{Conclusions}{\label{sec:concl}}
Once we know the 4D physics line we can do a simulation of the twisted
box with some convenient observable, and measure the mass through eq.~\ref{eq:twistav}. Care should be taken, as emphasized by Kajantie et al.~\cite{debye}, that we start in the symmetric phase and then move to the 4D physics line. In so doing we will stay on the physical branch of the hysteresis
curve for the mass, that we will meet when crossing the transition curve.
Nethertheless our discussion of the {\it location} of the critical curve underlines the importance to know wether the 4D physics line
lies for small x in the symmetric phase or in the broken phase.
\section*{Acknowledgments}
One of us (C.P.K.A.) thanks the organizers of this conference for their
hospitality and for the occasion to present this material.
\section*{References}
|
\section{Introduction}
The quantitative theory of stellar structure is more than 100 years
old {\Cse{Emd07}} and our understanding of the stellar interior has
improved dramatically during this time, especially since it became
possible to construct detailed stellar models with the help of
computers in the 1950's. However, even today, our understanding of
many observable properties of massive stars (${M_{\mathrm{ZAMS}}} \gtrsim 8 {\mathrm{M}_{\odot}}$,
$\log L/{\mathrm{L}_{\odot}} \gtrsim 4$) remains rudimentary.
Aside from comparatively minor uncertainties remaining in the
opacities and nuclear physics, the major frontiers in the study of
stars, and indeed stellar evolution in general, are proper treatments
of convection, mass loss, and rotation. This paper is the first in a
series concerning the effects of rotation and angular momentum
transport on the evolution of stars massive enough that a single one
can become a supernova ($M \gtrsim 8\,{\mathrm{M}_{\odot}}$).
The first to recognize the importance of rotation for celestial bodies
was Sir Isaac Newton. Early studies of rotating, self-gravitating,
incompressible fluids were carried out by McLaurin, Jakobi,
Poincar\'e, and Schwarzschild. Additional important contributions to
the numerical treatment of rotating stars were provided by
{\cite{KT70}}. {\cite{KMT70}} performed calculations taking these
effects into account using a simple model for angular momentum
transport. Studies with artificial rotation laws were carried out by
{\cite{ES76}}. In their pioneering work, {\cite{ES78}} considered
several rotationally induced instabilities, made order-of-magnitude
estimates for their efficiencies, and performed time-dependent stellar
evolution calculations of rotating massive stars up to the ignition of
carbon burning. Later, {\cite{pin89}} introduced a parameterization
of the poorly known efficiencies of the rotationally induced transport
processes of {\cite{ES78}} and gauged them to solar models. The
formalism we shall employ here is based largely upon these two works.
We differ, however, in using more recent data to calibrate the
uncertain efficiencies for angular momentum and composition transport
in this formalism and especially in following the stars past carbon
burning, all the way to the presupernova state.
Our formalism (discussed in detail, in $\S$2) is relatively simple
compared to others used in recent studies of rotation
during hydrogen and helium burning, e.g., {\cite{Lan92}},
{\cite{Den94}}, {\cite{ery94}}, {\cite{CZ92}}, {\cite{Zah92}},
{\cite{USS96}}, {\cite{tal97}}, {\cite{Meyn97}}, and {\cite{MZ98}},
but easier to understand and implement, and more easily extrapolable
to the late stages of stellar evolution. Indeed our poor understanding,
especially during the late stages of stellar evolution, of both convection
and possible modifications to angular momentum transport by magnetic
fields (not considered in the present work nor in the papers cited
above) suggests that it is worth trying something simple first.
Most of the rotation physics described in {\Sect{rotphys}} can already
be found, with slight modifications, in previous papers. However,
since this is the first in a series of papers, it will facilitate our
presentation to have all the relevant equations collected in one
place. We also correct several errors. In previous publications,
e.g. in the equation for the secular shear instability, and cast the
results in a consistent notation.
Following a summary of how we model various rotationally induced
instabilities (\Sect{rotphys}) and a discussion of the uncertain
parameters of the model (\Sect{calib}), we discuss the implementation
of this physics in the stellar codes in {\Sect{NumSolve}} and give an
overview of the initial models in {\Sect{InitialCond}}. The evolution
during hydrogen burning and helium burning is discussed in
{\Sects{cHburn}} and {\Sectff{cHeBurn}}, respectively. In
{\Sect{CmpOther}} we compare our results to the works of other
authors. The late evolution is discussed in {\Sect{LateEv}} and the
final angular momentum distribution at the presupernova stage is given
in {\Sect{j:preSN}}. Its implications are discussed in
{\Sect{youngPSR}} and a summary and our conclusions are given in
{\Sect{SumConcl}}.
Discussion of the details of observable parameters (evolution in the
HR diagram, surface abundances, lifetimes) and presupernova
nucleosynthesis are deferred to future papers.
\section{Rotation and Mixing in Massive Stars}
\lSect{rotphys}
\subsection{Modification to the stellar structure equations}
\lSect{ModRot}
In rotating stars, centrifugal forces act on the matter and lead to
deviations from spherical symmetry. For slow to moderate rotation
these deformations remain rotationally symmetric {\Cite{Tas78}}. Only
if the rotational energy exceeds a notable fraction of the binding
energy of the star does genuine triaxial deformation result.
In this work we consider only the case of ``slow'' rotation, i.e.,
where no triaxial deformations are expected. Some stars may reach
``critical'' rotation velocity ({\Sect{Omega}}) at their surfaces
during brief stages of their evolution {\Cite{HL98}}. However, except
for possibly modifying the mass loss rate ({\Sect{Omega}}), this
affects only the very outermost layers and is not expected to have a
big influence on the results of this paper.
Even for slow rotation, the shapes of surfaces of constant pressure,
constant density, and constant temperature are affected by the
centrifugal potential and thus deviate from spherical symmetry. The
momentum equation and the energy transport equation for spherically
symmetric stars must be modified to take this effect into account.
In this work, the centrifugal force is included following
{\cite{KT70}} in the approximation of {\cite{ES76}} and applied to the
hydrodynamic stellar structure equations {\Cite{Fli93}}. In this
approach, mass shells correspond to isobars instead of spherical
shells. Corrections are applied to the acceleration and the radiative
temperature gradient. According to {\cite{Zah75}}, {\cite{CZ92}}, and
{\cite{Zah92}}, anisotropic turbulence acts much stronger on isobars
than in the perpendicular direction. This enforces ``shellular''
rotation rather than cylindrical rotation {\Cite{MM97}}, and it sweeps
out compositional differences on isobars. Therefore it can be assumed
that matter on isobars is approximately chemically homogeneous.
Together with the shellular rotation this allows us to retain a
one-dimensional approximation. The specific angular momentum, $j$, of
a mass shell is treated as a local variable and the angular velocity,
$\omega$, is computed from the specific moment of inertia, $i$. The
time-dependent angular momentum redistribution is discussed in
{\Sect{AngTrans}}, and its influence on transport processes in
{\Sect{RotMix}}. We begin here by describing the modification to the
stellar structure equations of non-rotating stars {\Csa{ES76,MM97}}.
Let ${V_{\!\mathrm{P}}}$ be the volume enclosed by a surface of constant pressure,
$P$, and $\SP:=\partial {V_{\!\mathrm{P}}}$ its surface area. Then its
``radius'', ${r_{\!\mathrm{P}}}$, is defined as the radius of a sphere of the same
volume, ${V_{\!\mathrm{P}}}=4\pi {r_{\!\mathrm{P}}}^3/3$, and the equation of continuity becomes
\begin{equation}
\dxdycz{\mP}{{r_{\!\mathrm{P}}}}{t}=4\pi{r_{\!\mathrm{P}}}^2\rho
\;,
\lEq{mDefRot}
\end{equation}
where $\rho$ is the density and $\mP$ the mass enclosed by $\SP$.
For quantities varying on isobars, a mean value is
defined by
\begin{equation}
\av{\;\cdot\;} := \frac{1}{\SP} \oint_{\SP}\cdot\;\;{\mathrm d}\sigma\;,
\end{equation}
where ${\,\D\!\!\;}\sigma$ is an element of isobaric surface area. The
effective gravitational acceleration $\vec{g}$ is normal to $\SP$.
For the equation of momentum balance, one finds \Cite{ES76}
\begin{equation}
\dxdycz{P}{\mP}{t}=
-\frac{G \mP}{4\pi {r_{\!\mathrm{P}}}^4} {f_{\mathrm{P}}}
-\frac{1}{4\pi {r_{\!\mathrm{P}}}^2}\dxdycz{^2 {r_{\!\mathrm{P}}}}{t^2}{\mP}
\;,
\lEq{MomBalRot}
\end{equation}
where $G$ is the gravitational constant, $P$, the pressure, $t$, the
time, and the inertia term (last term) is added here. The influence
of rotation is described by the quantity ${f_{\mathrm{P}}}$
\begin{equation}
{f_{\mathrm{P}}} := \frac{4\pi {r_{\!\mathrm{P}}}^4}{G\mP\SP}
\av{g^{-1}}^{-1}
\;,
\end{equation}
where $g:=\abs{\vec{g}}$. The radiative temperature gradient then
takes the form
\begin{equation}
\dxdycz{\ln T}{\ln P}{t} =
\frac{3\kappa}{16\pi a c G}
\frac{P}{T^4}
\frac{\LP}{\mP}
\frac{{f_{\mathrm{T}}}}{{f_{\mathrm{P}}}}
\SBrak{1+\frac{{r_{\!\mathrm{P}}}^2}{G \mP {f_{\mathrm{P}}}}
\dxdycz{^2 {r_{\!\mathrm{P}}}}{t^2}{\mP}}^{\!-1}
\;,
\lEq{NabRadRot}
\end{equation}
where $\kappa$ is the opacity, $T$ the temperature, $a$ the radiation
constant, and $\LP$, the energy flux through $\SP$. The last
factor on the right hand side is included to account for inertia as it
follows from the momentum equation {\Cite{Fli93}}, and
\begin{equation}
{f_{\mathrm{T}}} := \Brak{\frac{4\pi {r_{\!\mathrm{P}}}^2}{\SP}}^{\!2}
\Brak{\av{g}\av{g^{-1}}}^{-1}
\;.
\end{equation}
For the derivation of these formulae and for a numerical evaluation of
${f_{\mathrm{T}}}$ and ${f_{\mathrm{P}}}$, see {\cite{ES76}}. The equations for ${f_{\mathrm{T}}}$ and ${f_{\mathrm{P}}}$
are solved iteratively with the stellar structure equations in order
to obtain consistent models {\Cite{ES76,Fli93}}. In the rest of this
work, the subscript $P$ is omitted (except for ${f_{\mathrm{P}}}$).
There is, in principle, an inconsistency between the assumption of
shellular rotation and the method described by {\cite{KT70}}, i.e.,
the assumption of shellular rotation does not generally lead to a
conservative potential as it does for a constant rotation rate on
cylinders, which is used by {\cite{KT70}}. However, {\cite{MM97}}
show that replacing the average $\av{\;\cdot\;}$ by ``appropriate mean
values'', i.e., reinterpreting the quantities describing the stellar
structure as the mean values over the isobars, allows one to keep the
formalism of {\cite{KT70}} as a good approximation.
\subsection{Ordinary mixing in the absence of rotation}
\lSect{MixInst}
Compositional mixing is generally treated as a diffusive process and
implemented by solving the diffusion equation
\begin{equation}
\dxdycz{X_n}{t}{m}=\dxdycz{}{m}{t} \SBrak{(4\pi r^2 \rho)^2 D
\dxdycz{X_n}{m}{t}}+\DxDyInd{X_n}{t}{\mathrm{nuc}}
\lEq{Diff}
\;,
\end{equation}
where $D$ is the diffusion coefficient constructed from the sum of
individual mixing processes and $X_n$, the mass fraction of species
$n$. The second term on the right hand side accounts for nuclear
reactions. At the inner and outer boundary reflecting conditions
are used:
\begin{equation}
\At{\dxdycz{X_n}{m}{t}}{m=0}=0=\At{\dxdycz{X_n}{m}{t}}{m=M(t)}
\;.
\end{equation}
Mixing, burning, and mass loss are treated as separate, sequential
operations. The different contributions to the diffusion coefficient,
$D$, are discussed in the following sections.
\subsubsection{Convection and overshooting}
\lSect{Conv}
\pFig{shear}
Convection occurs when the temperature gradient exceeds the adiabatic
condition, as modified by any gradient in mean molecular weight, $\mu$
(\Fig{shear}). That is, a stratification is stable against convection
if
\begin{equation}
{\nabla_{\!\mathrm{ad}}}-{\nabla}+\frac{\varphi}{\delta}{\nabla_{\!\mathrm{\mu}}} \geq 0
\lEq{Ledoux}
\end{equation}
{\Ce{KW91}}. This is the so-called Ledoux criterion for convection.
Here the common definitions are used:
\begin{equation}
{\nabla_{\!\mathrm{ad}}}:=\dxdycz{\,\ln T}{\,\ln P}{\mathrm{ad}}
\;,\quad
{\nabla_{\!\mathrm{\mu}}}:=\DxDy{\,\ln \mu}{\,\ln P}
\;,\quad
{\nabla}:=\DxDy{\,\ln T}{\,\ln P}
\;,
\end{equation}
\begin{equation}
\delta:=-\dxdycz{\ln\rho}{\ln T}{\mu,P}
\;,\qquad
\varphi:=\dxdycz{\ln\rho}{\ln\mu}{P,T}
\;.
\end{equation}
The index ``ad'' stands here for ``at constant entropy {\emph{and}}
composition''.
The diffusion coefficient for composition mixing is treated according
to the mixing-length theory {\Cite{Vit53,Bom58}}:
\begin{equation}
{D_{\mathrm{conv}}}:={\alpha_{\mathrm{MLT}}}\HP{v_{\mathrm{conv}}}/3
\;,
\end{equation}
where ${v_{\mathrm{conv}}}$ is the convective velocity. The pressure scale-height
is defined for the hydrostatic case by
\begin{equation}
\HP:=-\DxDy{r}{\ln P}
=
\frac{P}{\rho g}
\;.
\end{equation}
The local gravitational acceleration is given by $g=Gm/r^2$. In this
work a mixing-length parameter of ${\alpha_{\mathrm{MLT}}}=1.5$ {\Cite{Lan91B}} is used.
The mixing performed by convection is fast in comparison to most of
the other time-scales relevant for the stellar evolution. It operates
on the local dynamical time-scale and usually manages to smooth out
any compositional inhomogeneities in the regions where it is active.
Only when the time-scale of thermonuclear burning becomes comparable
to that of convection, as, e.g., during central silicon and shell
oxygen burning, can notable gradients persist.
In the present work ``overshooting'' of the convection into the
convectively stable regime defined by {\Eq{Ledoux}} is neglected. It
will be shown that rotation leads to mixing above the convective core
of massive stars during central hydrogen burning and thereby to the
formation of more massive helium cores later in the evolution. In
order to obtain such mixing, large overshooting is often introduced in
literature {\Ce{CS91,sch92}}, but moderate rotation can lead to
similar effects.
\subsubsection{Semiconvection}
\lSect{Semiconv}
Semiconvection is a secular instability which can occur in
non-rotating stars. According to a local, linear stability analysis by
{\cite{Kat66}}, it is an oscillatory instability which appears in
regions where an unstable temperature gradient is stabilized against
convection by a sufficiently large gradient in the mean molecular
weight ($\mu$-gradient), i.e., it lives in the regime
\begin{equation}
\frac{\varphi}{\delta}{\nabla_{\!\mathrm{\mu}}} \geq {\nabla}-{\nabla_{\!\mathrm{ad}}} \geq 0
\lEq{semiconv}
\end{equation}
{\CiteA{; and \Fig{shear}}{KW91}}. Heat transfer between a
displaced mass element and its surrounding causes the growth of the
instability on the local thermal time-scale.
In the code STERN (see {\Sect{STERN}}), semiconvection is treated
following {\cite{LSF83}}. The diffusion coefficient for this process
is computed from
\begin{equation}
{D_{\mathrm{sem}}}=\frac{{\alpha_{\mathrm{sem}}} K}{6{c_{\mathrm{P}}}\rho}
\frac{{\nabla}-{\nabla_{\!\mathrm{ad}}}}{{\nabla_{\!\mathrm{ad}}}-{\nabla}+\frac{\varphi}{\delta}{\nabla_{\!\mathrm{\mu}}}}
\;,\quad
K=\frac{4acT^3}{3\kappa\rho}
\;,
\lEq{Dsem+K}
\end{equation}
where $K$ is the thermal conductivity and ${c_{\mathrm{P}}}$ the specific heat
at constant pressure. As proposed by {\cite{Lan91B}}, an efficiency
parameter of ${\alpha_{\mathrm{sem}}}=0.04$ is adopted here.
In KEPLER (see {\Sect{KEPLER}}) semiconvection is computed from
{\Cite{WZW78,WW93}}
\begin{equation}
{{D'}_{\mathrm{\!\!sem}}}=\frac{1}{6}{\alpha_{\mathrm{MLT}}}^2{v_{\mathrm{sem}}}\HP
\;,
\end{equation}
where the velocity ${v_{\mathrm{sem}}}$ is determined through
\begin{equation}
{v_{\mathrm{sem}}}=\sqrt{\Brak{{\nabla}-{\nabla_{\!\mathrm{ad}}}}\frac{P\delta}{g\rho^2}\DxDy{P}{r}}
\;.
\end{equation}
The diffusion coefficient is limited to a fraction ${\alpha_{\mathrm{sem}}}$ of the
radiative diffusion coefficient
\begin{equation}
{D_{\mathrm{rad}}}=\frac{K}{\rho\cV}
\end{equation}
by means of
\begin{equation}
{D_{\mathrm{sem}}}=\frac{{\alpha_{\mathrm{sem}}}{D_{\mathrm{rad}}}{{D'}_{\mathrm{\!\!sem}}}}{{{D'}_{\mathrm{\!\!sem}}}+{\alpha_{\mathrm{sem}}}{D_{\mathrm{rad}}}}
\;.
\end{equation}
As usual, $\cV$ denotes the specific heat at constant volume. In
this work a value of ${\alpha_{\mathrm{sem}}}=10^{-4}$ is used in KEPLER, which results
in a comparable efficiency for semiconvection as the value used for
STERN {\Cite{Woo97:pc}}.
\subsection{Rotationally induced mixing}
\lSect{RotMix}
In this work, the mixing processes discussed in {\cite{ES78}} are
included in a parametric way, following the work of {\cite{pin89}}.
Five different processes are considered. To account for the uncertain
mixing efficiency of each, they are weighed by efficiency factors
{\CiteA{; \Sect{calib}}{pin89}} and then added to the diffusion
coefficient, $D$, in the diffusion equation {\Eqff{Diff}}.
\subsubsection{Dynamical shear instability}
\lSect{DSI}
\pFig{DSI}
Dynamical shear instability occurs when the energy that can be gained
from the shear flow becomes comparable to the work that has to be done
against the gravitational potential for the adiabatic turn-over of a
mass element (``eddy''). This means that it is stabilized by density
gradients. Since there is no work required to mix on isobars, this
instability can work very efficiently on those {\CITE{horizontal
turbulence; }{Zah92}} and thus enforce rigid rotation horizontally
{\Cite{ES78,pin89}}. Thus chemical inhomogeneities are smoothed on
isobars. This, together with the so called baroclinic instability,
which also acts barotropic for shear on isobars on a dynamical time
scale {\Cite{Zah83}}, justifies the assumption of shellular
rotation and that the composition is only a function of the isobars
(\Sect{ModRot}).
The linear condition for stability is given by
\begin{equation}
{R_{\mathrm{i}}}:=\frac{\rho\delta}{P}\Brak{{\nabla_{\!\mathrm{ad}}}-{\nabla}+\frac{\varphi}{\delta}{\nabla_{\!\mathrm{\mu}}}}
\Brak{g \DxDy{\ln r}{\omega}}^{\!2}>{R_{\mathrm{i,c}}}
\approx\frac{1}{4}
\lEq{def:ri}
\end{equation}
for the case of a rotating fluid {\Cite{Zah74}}. Here, $\omega$ is the
angular velocity, ${R_{\mathrm{i}}}$, Richardson number, and ${R_{\mathrm{i,c}}}$, its critical
value, about $1/4$. Note that the term for ${\nabla_{\!\mathrm{\mu}}}$ in {\Eq{def:ri}}
was omitted in the original work by {\cite{ES78}} and {\cite{pin89}}.
The corresponding diffusion coefficient is computed from the spatial
extent of the unstable region ${d_{\mathrm{inst}}}$, limited to a pressure
scale-height, and the local dynamical time-scale {\Csa{ES78}}:
\begin{equation}
{D_{\mathrm{DSI}}}=\SBrak{\Min{{d_{\mathrm{inst}}},\HP}
\Brak{1-\Max{\frac{{R_{\mathrm{i}}}}{{R_{\mathrm{i,c}}}},0}}}^2
/{\tau_{\mathrm{dyn}}}
\;,
\end{equation}
where the dynamical time-scale is defined by
\begin{equation}
{\tau_{\mathrm{dyn}}}:=\sqrt{r^3/(G\,m)}
\;.
\end{equation}
Furthermore, it is assumed that the instability is weaker when the
deviation from the Richardson criterion is smaller. This is accounted
for by a factor $(1-{R_{\mathrm{i,c}}}/{R_{\mathrm{i}}})^2$, which is limited to the range
$[0,1]$. For ${R_{\mathrm{i}}} > {R_{\mathrm{i,c}}}$, the flow is assumed to be stable against
the dynamical shear instability and thus ${D_{\mathrm{DSI}}}$ is set to $0$.
\subsubsection{Solberg-H{\o}iland instability}
\pFig{SHI}
The Solberg-H{\o}iland instability arises if an adiabatically
displaced mass element experiences a net force (the sum of gravity,
buoyancy and centrifugal force) that has components in the direction
of the displacement only. {\cite{Was46}} gives a condition for the
stability against axisymmetric adiabatic perturbations of this kind.
It separates into two scalar conditions. At the equator the condition
for stability in the vertical direction is
\begin{equation}
{R_{\mathrm{SH}}}:=\frac{g}{\rho}\SBrak{\Brak{\DxDy{\rho}{r}}_{\!\mathrm{ad}}
-\DxDy{\rho}{r}}
+\frac{1}{r^3}\DxDy{}{r}\Brak{r^2\omega}^2 \geq 0
\lEq{SHcA}
\end{equation}
{\Cite{Tas78,ES78}}. If the specific angular momentum $j\sim r^2\omega$
is constant with $r$, the last term on the left-hand side vanishes and
the Ledoux criterion results --- not the Schwarzschild criterion as
stated by \cite{ES78}. This can be seen by rewriting the condition
for stability
\begin{equation}
{R_{\mathrm{SH}}}:=\frac{g\delta}{\HP}\SBrak{{\nabla_{\!\mathrm{ad}}}-{\nabla}+\frac{\varphi}{\delta}{\nabla_{\!\mathrm{\mu}}}}
+\frac{1}{r^3}\DxDy{}{r}\Brak{r^2\omega}^2 \geq 0
\lEq{SHcB}
\end{equation}
and comparing it with \Eq{Ledoux}. If, on the other hand, the medium
is marginally stable to convection, the first term on the right
hand side vanishes and the Rayleigh criterion results {\Cite{Tas78,KW91}}.
Note that this instability only occurs in regions of
{\emph{decreasing}} specific angular momentum (\Fig{SHI}) and is
strongly suppressed in stable stratifications
(${\nabla}<{\nabla_{\!\mathrm{ad}}}+\frac{\varphi}{\delta}{\nabla_{\!\mathrm{\mu}}}$).
The diffusion coefficient resulting from the Solberg-H{\o}iland
instability is estimated in a way similar to that for the dynamical
shear instability. The extent of the unstable region, ${d_{\mathrm{inst}}}$,
limited to the pressure scale-height, is used as the characteristic
length-scale, and the dynamical time-scale is used as characteristic
time-scale:
\begin{equation}
{D_{\mathrm{SHI}}}=\SBrak{\Min{{d_{\mathrm{inst}}},\HP} \Brak{\frac{r\,{R_{\mathrm{SH}}}}{g}}}^2/{\tau_{\mathrm{dyn}}}
\;.
\end{equation}
Again, as for the dynamical shear instability, a factor of order unity
($r\,{R_{\mathrm{SH}}}/g$) was introduced to smoothly turn on the instability as
the criterion for stability gets increasingly violated, and ${D_{\mathrm{SHI}}}$ is
set to $0$ wherever the stability criterion is fulfilled.
\subsubsection{Secular shear instability}
\lSect{SSI}
The strict criterion for dynamical shear instability can be relaxed
considerably by allowing for thermal adjustment of radial
perturbations. However, this process then operates only on a thermal
time-scale, and is therefore a secular process. Gradients in the
mean molecular weight, which may inhibit the occurrence of the
instability, also have to be taken into account.
According to {\cite{ES78}}, the following two conditions have to be
violated simultaneously for this instability to set in (\Fig{shear}):
\begin{equation}
{R_{\mathrm{is,1}}}:=\frac{{{\mathcal{P}_{\!\mathrm{r}}}}\,{R_{\mathrm{e,c}}}}{8}\frac{\rho\delta}{P}\Brak{{\nabla_{\!\mathrm{ad}}}-\nabla}
\Brak{g\DxDy{\ln r}{\omega}}^{\!2}
>{R_{\mathrm{i,c}}}
\end{equation}
{\Cite{Tow58,Zah75}} because of the relaxed condition for the
temperature gradient, and
\begin{equation}
{R_{\mathrm{is,2}}}:=\frac{\rho\varphi{\nabla_{\!\mathrm{\mu}}}}{P}
\Brak{g\DxDy{\ln r}{\omega}}^{\!2}
>{R_{\mathrm{i,c}}}
\lEq{SSIb}
\end{equation}
since the condition for the $\mu$-gradient is not relaxed. The latter
formula follows from the physical arguments of {\cite{ES78}}, but
corrects an error in their Eq.~(10). For the critical Reynolds
number, ${R_{\mathrm{e,c}}}$, a value of $2500$ is assumed in this work {\CITE{but
see also}{RZ99}}. The Prandtl number, ${{\mathcal{P}_{\!\mathrm{r}}}}$, is defined as the ratio
of the thermal diffusion time-scale to the angular momentum diffusion
time-scale, and is estimated according to {\cite{Tas78}}:
\begin{equation}
{{\mathcal{P}_{\!\mathrm{r}}}}=
\frac{\cV\Brak{{\mu_{\mathrm{p}}}+{\mu_{\mathrm{r}}}}}{\chi}
\;,
\end{equation}
where the coefficients of shear viscosity of the plasma and by
radiation are computed according to
\begin{equation}
{\mu_{\mathrm{p}}}\approx0.406\,\frac{\sqrt{\mi (\kB T)^5}}
{\Brak{{Z_{\mathrm{i}}} e}^{4}\ln\Lambda}
\;,\qquad
{\mu_{\mathrm{r}}}=\frac{4 a T^4}{15 c \kappa \rho}
\end{equation}
{\Cite{Spi62,Tas78}}, respectively. The quantity $\Lambda$ is the
ratio of the cut-off length for ion collisions, which is taken as the
ratio of the Debye length, to the impact parameter for a $\pi/2$
deflection for Rutherford scattering of the ions, i.e.,
\begin{equation}
\Lambda=\frac{2}{3 e^3}\sqrt{
\frac{\mi \Brak{\kB T}^{3}}{\pi \rho {Z_{\mathrm{i}}}^5}}
\end{equation}
\CITE{for details, see}{Spi62}. Here, $e$ is the charge of the
electron in e.s.u., $c$ the velocity of light, $\kB$ Boltzmann's
constant, ${Z_{\mathrm{i}}}$ the charge number of the ion, and $\mi$ its mass. It
should be noted that for burning phases beyond hydrogen burning, as
well as for helium, carbon, or oxygen stars, it is important to take
the ${Z_{\mathrm{i}}}$-dependence of the plasma viscosity into account. The
quantity $\Lambda$ enters only logarithmically and $\ln \Lambda$ is
$\sim25$. At temperatures below the Fermi temperature, depending
somewhat on the chemical composition, the ion viscosity dominates over
the electron contribution. For the evaluation of the formulae above,
complete ionization is assumed.
If magnetic fields and neutrinos are neglected, the thermal
conductivity is given by $\chi\approx K$ {\CITE{\Eq{Dsem+K};
}{Tas78}}. The opacity, $\kappa$, used in this work takes into
account the energy transport by radiation as well as heat conduction
by degenerate electrons. Following {\cite{ES78}}, the circulation
velocity associated with this process is computed from the time-scale
and the length-scale of the turbulent elements,
\begin{equation}
{v_{\mathrm{SSI}}}=\sqrt{\frac{\nu}{{R_{\mathrm{e,c}}}}\DxDy{\omega}{\ln r}}
\;,
\end{equation}
limited to the adiabatic sound velocity, ${c_{\mathrm{s}}}$. The kinematic
viscosity, $\nu$, is given by {\Cite{Tas78}}
\begin{equation}
\nu=\frac{{\mu_{\mathrm{p}}}+{\mu_{\mathrm{r}}}}{\rho}
\;.
\end{equation}
For the typical length-scale the velocity scale height of the flow is
assumed,
\begin{equation}
{H_{\mathrm{v,SSI}}}:=\abs{\DxDy{r}{\ln {v_{\mathrm{SSI}}}}}
\;,
\end{equation}
limited to the pressure scale height. The resulting diffusion
coefficient is given by
\begin{equation}
{D_{\mathrm{SSI}}}=\Min{{v_{\mathrm{SSI}}},{c_{\mathrm{s}}}}
\Min{{H_{\mathrm{v,SSI}}},\HP}
\Brak{1-\frac{\Max{{R_{\mathrm{is,1}}},{R_{\mathrm{is,2}}}}}{{R_{\mathrm{i,c}}}}}^{\!2}
\;.
\end{equation}
Again, the instability is smoothly turned on with increasing
violation of the stability criteria (term in the last bracket).
In recent work, {\cite{MM96}},{\cite{Mae97}}, and {\cite{MZ98}}
reconsidered the interaction of thermal diffusivity, horizontal
turbulence (due to the baroclinic instability), and vertical shear.
An important conclusion that can be drawn from their work is that
$\mu$-gradients might not completely suppress the occurrence of the
shear instability, since the medium is already turbulent due to the
baroclinic instability. Consequently, some mixing can occur
{\Cite{Mae97}}. In the present work, we parameterize the efficiency
of the secular shear instability for chemical mixing and of the
$\mu$-gradients in suppressing its occurrence (\Sect{uncertMix}).
\subsubsection{Eddington-Sweet circulation}
\lSect{ES}
As first shown by {\cite{Zei24A,Zei24B}} for rigid rotation, and later
by {\cite{BK59}} for a general rotation law, a rotating star cannot be
in hydrostatic and radiative thermal equilibrium at the same time.
This is so because surfaces of constant temperature and constant
pressure do not coincide. Consequently, large-scale circulations
develop. Since inhomogeneities on isobars are quickly smoothed out by
the horizontal turbulence only the perpendicular ($\approx$ radial)
component of the circulation velocity is considered here, and the
process is approximated by diffusion along the radial coordinate.
\cite{Kip74} estimated the circulation velocity as
\begin{equation}
{v_{\mathrm{e}}}:=\frac{{\nabla_{\!\mathrm{ad}}}}
{\delta\,\Brak{{\nabla_{\!\mathrm{ad}}}-\nabla}}
\frac{\omega^2 r^3 l}{\Brak{G m}^2}
\Brak{\frac{2 \Brak{{\varepsilon_{\mathrm{n}}}+{\varepsilon_{\mathrm{\nu}}}} r^2 }{l}
-\frac{2 r^2}{m}
-\frac{3}{4 \pi \rho r}}
\;.
\lEq{vESO}
\end{equation}
In the presence of $\mu$-gradients, meridional circulation has to
work against the potential and thus might be inhibited or suppressed
{\Cite{Mes52,Mes53}}. Formally, this can be written as a ``stabilizing''
circulation velocity,
\begin{equation}
{v_{\mathrm{\mu}}}:=\frac{\HP}{{\tau^*_{\mathrm{KH}}}}
\frac{\varphi{\nabla_{\!\mathrm{\mu}}}}{\delta\,\Brak{\nabla-{\nabla_{\!\mathrm{ad}}}}}
\lEq{vmu}
\end{equation}
{\Cite{Kip74,pin89}}, where
\begin{equation}
{\tau^*_{\mathrm{KH}}}:=\frac{G m^2}{r \Brak{l - m {\varepsilon_{\mathrm{\nu}}}}}
\lEq{tauKHx}
\end{equation}
is the local Kelvin-Helmholtz time-scale, used here as an estimate for
the local thermal adjustment time-scale of the currents
{\Cite{pin89}}. The spatial extent of the currents is typically of
the order of the radius coordinate $r$. Here, neutrino losses are
taken into account, because they reduce the thermal time-scale in the
late stages of the stellar evolution significantly. Note that
${\varepsilon_{\mathrm{\nu}}}$ is defined as the energy {\emph{generation}} rate due to
neutrino losses and therefore is negative. This increases the
numerator in the definition of the local Kelvin-Helmholtz time-scale
and thus decreases ${\tau^*_{\mathrm{KH}}}$.
For the evaluation of the diffusion coefficient, the sign of the
circulation velocity does not matter, but the stabilizing ``currents''
due to $\mu$-gradients always point in the direction opposite to the
meridional flow, thus resulting in a reduction of the effective
circulation velocity. The velocity is then computed from
\begin{equation}
{v_{\mathrm{ES}}}:=\Max{\abs{{v_{\mathrm{e}}}}-\abs{{v_{\mathrm{\mu}}}},0}
\lEq{vES}
\end{equation}
{\CiteA{; and {\Fig{GSF}}}{ES78}}. The diffusion coefficient is
calculated as the the product of the circulation velocity and a
typical length-scale for the circulation. This is assumed to be the
minimum of the extent ${d_{\mathrm{inst}}}$ of the instability and the velocity
scale-height
\begin{equation}
{H_{\mathrm{v,ES}}}:=\abs{\DxDy{r}{\ln {v_{\mathrm{ES}}}}}
\end{equation}
{\Cite{ES78}}, i.e.,
\begin{equation}
{D_{\mathrm{ES}}}:=\Min{{d_{\mathrm{inst}}},{H_{\mathrm{v,ES}}}}{v_{\mathrm{ES}}}
\;.
\end{equation}
In recent work, {\cite{CZ92,Zah92,USS96,tal97,MZ98}} have discussed
several improvements to the theory of meridional circulation and its
interaction with the baroclinic instability. In contrast to the
present work, their method requires the solution of a fourth order
differential equation in $\omega$, which is numerically very involved.
So far this method has only been used to investigate main sequence
stars. An interesting result of these work for the Eddington-Sweet
circulation is that the stabilizing effect of $\mu$-gradients,
entering through ${v_{\mathrm{e}}}$ in ${v_{\mathrm{ES}}}$ ({\Eqs{vESO}} and {\Eqff{vES}}),
may be reduced (\Sect{uncertMix}). The second important change to the
above estimate is that the interaction of the baroclinic instability
and the large-scale meridional reduces the mixing efficiency of the
Eddington-Sweet circulation in agreement with the numerical studies by
{\cite{pin89}} for the sun. We consider these effects when we perform
an empirical calibration of the mixing efficiencies in {\Sect{calib}}.
\subsubsection{Goldreich-Schubert-Fricke instability}
\lSect{GSF}
\pFig{GSF}
{\cite{GS67}} and {\cite{Fri68}} performed an analysis of stability
against axisymmetric perturbations (GSF instability). For the
inviscid limit {\mbox{(${{\mathcal{P}_{\!\mathrm{r}}}} \ll 1$)}}, which can be well assumed in
the interior of stars, they derive two conditions for stability in
chemically homogeneous stars {\Cite{Kip69}}:
\begin{equation}
\dxdy{j}{r} \geq 0
\qquad\mbox{and}\qquad
\dxdy{\omega}{z} = 0
\;.
\lEq{GSFc}
\end{equation}
The first condition is the secular analogue to the Solberg-H{\o}iland
stability criterion {\Eq{SHcB}}, where the stabilization by the
temperature gradient is removed due to thermal conduction. This is
similar to the relation between the secular and the dynamical shear
instability. The second condition in {\Eq{GSFc}} is the analogue to
the Taylor-Proudman theorem for slowly rotating incompressible fluids
{\Cite{Kip74,Tas78}}. If the rotational velocity depends on the
distance from the equatorial plane, i.e., the rotation profile is not
conservative, meridional flows will be driven. Also in this case, the
buoyancy force, which acts to suppress the instability, can be removed
by heat conduction. However, this occurs only on a thermal
time-scale. Interestingly, the typical velocities for both the
above processes are quite similar {\Cite{Kip74}}.
Since the second condition of {\Eq{GSFc}} is in general in
contradiction with the shellular rotation law enforced by the
baroclinic instability, except for the case of solid body rotation,
the GSF instability will tend to enforce uniform rotation in
chemically homogeneous regions {\Cite{ES78}}.
The dependence of the GSF instability on differential rotation is
stronger than that of Eddington-Sweet circulation, and the large-scale
circulation velocity in the equatorial plane can be estimated by
\begin{equation}
{v_{\mathrm{g}}}=\frac{2 {H_{\mathrm{T}}} r}{{H_{\mathrm{j}}}^2}
\Brak{1+2\DxDy{\ln r}{\ln \omega}}^{-1}
{v_{\mathrm{e}}}
=\frac{2{H_{\mathrm{T}}}}{{H_{\mathrm{j}}}}\,\DxDy{\ln \omega}{\ln r}
{v_{\mathrm{e}}}
\end{equation}
{\Cite{ES78,JK70,JK71}}. Here ${H_{\mathrm{T}}}:=-\Frac{{\mathrm d} r}{{\mathrm d} \ln T}$ is the
temperature scale-height and ${H_{\mathrm{j}}}:=\Frac{{\mathrm d} r}{{\mathrm d} \ln j}$, the
scale-height of the angular momentum distribution. The GSF
instability has the same $\mu$-dependence as Eddington-Sweet
circulation {\Cite{ES78}} and therefore the resulting circulation
velocity is computed in the same way, taking the stabilizing effect of
the $\mu$-gradient into account:
\begin{equation}
{v_{\mathrm{GSF}}}:=\Max{\abs{{v_{\mathrm{g}}}}-\abs{{v_{\mathrm{\mu}}}},0}
\;.
\end{equation}
Again, the diffusion coefficient is determined from the circulation
velocity, ${v_{\mathrm{GSF}}}$, and the minimum of the circulation velocity scale
height, ${H_{\mathrm{v,GSF}}}$, and the extent, ${d_{\mathrm{inst}}}$, of the instability:
\begin{equation}
{D_{\mathrm{GSF}}}:=\Min{{d_{\mathrm{inst}}},{H_{\mathrm{v,GSF}}}}{v_{\mathrm{GSF}}}
\;,
\end{equation}
where we define in the same way as above
\begin{equation}
{H_{\mathrm{v,GSF}}}:=\abs{\DxDy{r}{\ln {v_{\mathrm{GSF}}}}}
\;.
\end{equation}
{\Fig{GSF}} compares the parameter space in which the GSF and the
Eddington-Sweet instability operate. For small angular velocity
gradients the Eddington-Sweet circulation dominates, while the GSF
instability becomes more important as the differential rotation
increases. Note that for strong differential rotation the shear
instability also occurs (cf. {\Fig{shear}})
\subsection{Other instabilities}
\lSect{OtherInst}
The five instabilities discussed in the previous section are not
a complete list of all rotationally induced instabilities for
massive stellar evolution. However, they appear to be the most
relevant ones, or at least the best understood.
For the ABCD-instability {\Cite{SKR84}} and the triply diffusive
instability {\Cite{KS83}}, no reliable estimates of efficiency
exist. Furthermore, non-axisymmetric instabilities may also exist,
but are poorly investigated so far.
Another important issue is the interaction of the different
instabilities, and the interaction of rotation and rotationally
induced instabilities with the instabilities listed in
{\Sect{MixInst}}. The interaction of the shear instabilities and the
Eddington-Sweet circulation has been investigated by, e.g.,
{\cite{CZ92,Zah92,USS96,MM97,Mae97,TZ97,tal97,MZ98}}, and
semiconvection has recently also been included by {\cite{Mae97}} and
{\cite{MZ98}}. However, the effects of the interactions are not
large and therefore not taken into account in the present work.
Perhapes most importantly, we have neglected magnetic fields.
Magnetic fields might transport angular momentum by torques
{\CITE{$\sim r^3 B_{r}B_{\phi}$;}{Spr97}}, or cause instabilities by
magnetic buoyancy resulting from the winding up of magnetic field
lines by differential rotation. This could be effective even if the
initial field strength is small {\Cite{Spr97,SP98}}. Unfortunately,
little is known about either the strength of the initial field or the
efficiency of instabilities in amplifying the magnetic field. The
Velikhov-Chandrasekhar instability depends only on the presence of
magnetic fields, not on their strength, but it is efficiently
suppressed by $\mu$-gradients {\Cite{Ach78,Spr97}}. Detailed studies
of the action of magnetic fields inside stars must be left to future
investigations.
\subsection{Angular momentum transport}
\lSect{AngTrans}
Following {\cite{ES78}} and {\cite{pin89}}, we formulate the transport
of angular momentum as a diffusive process,
\begin{equation}
\dxdycz{\omega}{t}{m}=\frac{1}{i}\dxdycz{}{m}{t}
\SBrak{(4\pi r^2 \rho)^2 i\nu \dxdycz{\omega}{m}{t}}
- \frac{2\omega}{r}\dxdycz{r}{t}{m}
\Brak{\frac{1}{2}\DxDy{\,\ln i}{\,\ln r}}
\lEq{AngDiff}
\end{equation}
{\Cite{ES78}}, where $\nu$ is the turbulent viscosity and $i$, the
specific angular momentum of a shell at mass coordinate $m$. For a
spherical shell of constant density, inner radius $\ri$ and outer
radius ${r_{\!\mathrm{o}}}$, the specific moment of inertia, $i$, is given by
$i=0.4\,\Frac{{r_{\!\mathrm{o}}}^5-\ri^5}{{r_{\!\mathrm{o}}}^3+\ri^3}$; for a thin shell
of radius $r$ this simplifies to $i=2/3\,r^2$. The last term in
{\Eq{AngDiff}}, an advection term, accounts for contraction or
expansion of the layers at constant mass coordinate. The factor in
the last bracket on the right hand side vanishes if the gyration
constant $k:=i/r^2$ does not depend on $r$.
{\Eq{AngDiff}} is essentially a diffusion equation for $\omega$
along the ``moment of inertia coordinate'',
\begin{equation}
I(m):=\int_0^m i(m'){\,\D\!\!\;} m'
\;,
\end{equation}
defined analogously to the mass coordinate {\Ce{KW91}}. This equation
conserves angular momentum and leads to rigid rotation in a region of
extent $\ell$ whenever the diffusion time-scale, ${\tau_{\mathrm{D}}} :=
\ell^2/\nu$, is short in comparison to structural changes of the star.
Since the Eddington-Sweet circulation may redistribute angular
momentum by advection rather than by viscous stress {\Cite{Zah92}},
the equilibrium solution might deviate from rigid rotation assumed
here in regions where it is the dominant process. However, for
consistency to {\cite{ES78,pin89}} and for simplification of the
numerical treatment we stick with the prescription outlined above.
Compared to {\cite{TZ97}} we get very similar results at the end of
central hydrogen burning (see {\Sect{CmpOther}}).
At the inner and outer boundary, reflecting conditions similar to those
given in {\Eq{Diff}} for the compositional mixing are used.
At the surface of the star, the angular momentum contained
in the layers which are lost due to stellar winds is removed from
the star (\Sect{AngLoss}).
The turbulent viscosity, $\nu$, is determined as the sum of the
convective and semiconvective diffusion coefficients, and those from
rotationally induced instabilities {\CiteA{; {\Sect{calib}}}{ES78}}.
In contrast to {\cite{ES78}} and {\cite{pin89}}, in the present work the
transport equation for angular momentum is solved for the entire star
as a whole.
Since the evolutionary time-scale of the star is in most cases much
longer than the convective time-scale, {\Eq{AngDiff}} results in rigid
rotation in those regions. Unlike composition, which can show
significant gradients even inside convective regions due to burning
(e.g., during central silicon burning), angular momentum is locally
conserved, and therefore convective regions can more easily reach
rigid rotation than chemical homogeneity during hydrostatic burning
phases. This, however, does not hold if the respective layers are
contracting or expanding rapidly.
The approximation that convection leads to rigid rotation rather than
constant specific angular momentum seems to be justified, at least if
the rotational period is long in comparison to the convective time
scale, and it may also hold for more rapid rotation if convective
blobs can be assumed to scatter elastically {\Cite{KNL95}}. The
latitudinally averaged rotation rate of the solar convection zone
deviates from solid body rotation by less than $5\,\%$ {\Ce{ABC97}}.
\subsection{Enhanced mass loss due to rotation}
\lSect{MassLoss}
\lSect{Omega}
Mass loss from the stellar surface (``stellar winds'') significantly
affects the evolution of massive stars {\Cite{CM86}}. In the present
work, the empirical mass loss rate of {\cite{NJ90}} is used. For
Wolf-Rayet stars, the prescription of {\cite{Lan89B}} is applied. The
uncertainties in these mass loss rates are considerable due to the
uncertainties in the observational data and their interpretation.
These mass loss rates are further modified to account for the effect
of stellar rotation according to {\cite{FA86}}
\begin{equation}
{\dot{M}}(\omega) := {\dot{M}}(\omega=0) \times
\left(\frac{1}{1-\Omega}\right)^{\xi}
\;,\qquad
\xi\approx0.43
\lEq{MassLossRot}
\;,
\end{equation}
where
\begin{equation}
\Omega:=\frac{v}{{v_{\mathrm{crit}}}}
\;,
\lEq{Omega}
\end{equation}
is the ratio of the equatorial surface rotation rate to the critical
rotation rate defined by
\begin{equation}
{v_{\mathrm{crit}}}^2:=\frac{Gm}{r}\Brak{1-\Gamma}
\;.
\lEq{vcrit}
\end{equation}
The Eddington factor,
\begin{equation}
\Gamma:=\frac{\kappa L}{4 \pi c G m}
\;,
\lEq{Gamma}
\end{equation}
is evaluated only in the radiative part of the optical depth range
$\tau \in [2/3 , 100]$ {\Cite{Lam93,Lan97:LBV}}, where
$\tau(r)=\int_r^{\infty}\kappa\rho{\,\D\!\!\;} r$ has the usual definition.
The quantitative result for the $\Omega$-dependence of the mass loss
rate obtained by {\cite{FA86}} was questioned by {\cite{OCG96}}, who
performed hydrodynamic simulations of the winds of rotating hot stars
including the effect of non-radial radiation forces and
gravity-darkening in the approximation of {\cite{Zei24A,Zei24B}}.
In any case, the latitude dependence of the surface properties
(temperature, radiation flux, etc.) of rapidly rotating luminous stars
is largely unknown as {\cite{Kip77}} showed in a generalization of the
von Zeipel theorem that they depend strongly on the details of the
internal rotation law {\Csa{Mae99}}. However, the only crucial
ingredient for our model calculations, which is confirmed by
{\cite{OG97}}, is the fact that the latitudinally integrated mass loss
rate
increases strongly as the star approaches the $\Omega$-limit, so that
the star cannot exceed critical rotation, but rather loses more mass
and angular momentum {\Cite{Lan98}}.
\subsection{Angular momentum loss}
\lSect{AngLoss}
The loss of angular momentum from the surface due to stellar winds is
approximated by removing of the angular momentum along with the
surface layer, i.e.,
\begin{equation}
{\dot{J}}={\dot{M}}{j_{\mathrm{surf}}}
\;,
\end{equation}
where ${j_{\mathrm{surf}}}$ is the latitudinally averaged
specific angular momentum at the surface
of the star when the mass loss is assumed independent of latitude.
\section{Calibration of the mixing efficiencies}
\lSect{uncertMix}
\lSect{calib}
The diffusion coefficients used in this work are subject to
considerable uncertainties, as they result from order-of-magnitude
estimates of some of the relevant time- and length-scales. Therefore,
efficiency factors of order unity are introduced, in order to
calibrate the diffusion coefficients with observational data. This is
similar to the treatment of {\cite{pin89}}.
The first adjustable parameter is the ratio of
the turbulent viscosity to the diffusion
coefficient, $\fc:=D/\nu$. The contribution of the rotationally
induced instabilities to the diffusion coefficient is assumed to be
reduced by the factor $\fc$, while their full value enters the
turbulent viscosity,
\begin{equation}
D={D_{\mathrm{conv}}}+{D_{\mathrm{sem}}}+\fc\Brak{{D_{\mathrm{DSI}}}+{D_{\mathrm{SHI}}}+{D_{\mathrm{SSI}}}+{D_{\mathrm{ES}}}+{D_{\mathrm{GSF}}}}
\;,
\end{equation}
\begin{equation}
\nu={D_{\mathrm{conv}}}+{D_{\mathrm{sem}}}+{D_{\mathrm{DSI}}}+{D_{\mathrm{SHI}}}+{D_{\mathrm{SSI}}}+{D_{\mathrm{ES}}}+{D_{\mathrm{GSF}}}
\;.
\end{equation}
The second parameter, ${f_{\!\mathrm{\mu}}}\in[0,1]$, describes the sensitivity of the
rotationally induced mixing to $\mu$-gradients, i.e., ${\nabla_{\!\mathrm{\mu}}}$
is replaced by ${f_{\!\mathrm{\mu}}}{\nabla_{\!\mathrm{\mu}}}$.
In order to reproduce the surface {\I{7}{Li}} abundance in the sun,
{\cite{pin89}} introduced the factor $\fc\in[0,1]$. They found a value
of $\fc=0.046$ for their best fit. From theoretical work {\cite{CZ92}}
found a similar value, $\fc=1/30$, for the combined action of shear and
meridional circulation. This is the value chosen for most of the
models presented in this work (cf. {\Tab{InitModels}}).
\pFig{M-CNOHe-gauge}
\pFig{fc-gauge}
The best observational probe of rotationally induced mixing in
stars is the evolution of the
surface composition during central hydrogen burning. While lithium
and boron are depleted early during this phase {\Cite{VLL96,FLV96}},
since they are destroyed at relatively low temperatures, {\I{14}{N}},
is only produced at higher temperature, i.e., much deeper
inside the star. Therefore, an increase of nitrogen at the surface
should be accompanied by an decrease of carbon ({\I{12}{C}}) or, in
the case of even deeper mixing, oxygen ({\I{16}{O}}), which is
destroyed at even higher temperatures.
An enrichment of nitrogen of order $2\ldots3$ is observed for evolved
stars of about $10$ to $20\,{\mathrm{M}_{\odot}}$ {\Cite{GL92,Her94,vra98}}. Since
observations can only give the projected rotation rate and are also
restricted to low projected rotational velocities (\Cite{GL92,vra98}),
only a qualitative comparison with our models is possible.
The processing of carbon to nitrogen which occurs at core hydrogen
ignition does not introduce large $\mu$-gradients. Therefore, the
occurrence of a surface nitrogen enrichment and carbon depletion is
rather insensitive to ${f_{\!\mathrm{\mu}}}$. In contrast, any enrichment of helium
in O stars {\Cite{her92,HVM98}} strongly restricts ${f_{\!\mathrm{\mu}}}$.
Unfortunately, helium abundances are hard to measure and
correspondingly uncertain {\Cite{Her94}}.
For purposes of calibration, we computed evolutionary sequences for
solar metallicity stars in the mass range $4\,{\mathrm{M}_{\odot}}$ to $60\,{\mathrm{M}_{\odot}}$
through core hydrogen burning, adopting a typical zero-age main
sequence rotational velocity of $\sim 200\,{\km\,\Sec^{-1}}$
{\Cite{Sle70,Fuk82,Lang91,Hal96,Pen96}}. {\Fig{M-CNOHe-gauge}} shows
the surface values of helium, carbon, nitrogen, and oxygen at core
hydrogen exhaustion as function of the initial stellar mass for
various combinations of ${f_{\!\mathrm{\mu}}}$ and $\fc$.
A value of ${f_{\!\mathrm{\mu}}}=0.05$ reproduces an
enhancement of nitrogen by a factor of $2$ to $3$ in the mass range
$10\,{\mathrm{M}_{\odot}}$ to $20\,{\mathrm{M}_{\odot}}$, and results in a surface helium mass
fraction of $\sim40\,\%$ for the $60\,{\mathrm{M}_{\odot}}$ star, while the
enrichment remains quite small for stars below $20\,{\mathrm{M}_{\odot}}$. For
${f_{\!\mathrm{\mu}}}=0.01$, nitrogen and helium are clearly enriched too much for
stars below $30\,{\mathrm{M}_{\odot}}$. On the other hand, the nitrogen enrichment
might be too low for values of ${f_{\!\mathrm{\mu}}}\ge0.1$. Certainly, for
${f_{\!\mathrm{\mu}}}=0.25$ and ${f_{\!\mathrm{\mu}}}=1.0$ the nitrogen abundance for the most massive
stars ($30\,{\mathrm{M}_{\odot}}\ldots60\,{\mathrm{M}_{\odot}}$) is inconsistent with the
observations. The same is true for the helium abundances.
\pTab{InitModels}
In summary, ${f_{\!\mathrm{\mu}}}=0.05$ seems to be the best value (provided
$\fc=1/30$; see above). This set of parameters is used in the present
work for the models whose name ends with ``{\Mod{B}}''
(\Tab{InitModels}). The consequences of a variation of $\fc$ (for
fixed ${f_{\!\mathrm{\mu}}}=0.05$) is shown in {\Fig{fc-gauge}} for a $12\,{\mathrm{M}_{\odot}}$
star. For small values of $\fc$ the nitrogen abundance is too low,
while for large values, helium becomes quite high.
As discussed above, too much surface enrichment occurs with $\fc=1/30$
for small values of ${f_{\!\mathrm{\mu}}}$ ($\simle0.01$). Nevertheless, it is
interesting to investigate the case where $\mu$-gradients are
completely neglected, since the calibration of $\fc$ and ${f_{\!\mathrm{\mu}}}$ is not
unambiguous, and different combinations might result in similar
surface enrichments. The surface abundance, however, are the only
clear observational constraint, while the degree of internal mixing is
not directly observable. Therefore, a second parameter set of
$\fc=0.01$ and ${f_{\!\mathrm{\mu}}}=0$ is also used. The resulting surface
abundances (displayed as thick grey line in {\Fig{M-CNOHe-gauge}})
show quite similar enrichments. Models with this choice of $\fc$ and
${f_{\!\mathrm{\mu}}}$ do {\emph{not}} carry a ``{\Mod{B}}'' at the end of their name
(\Tab{InitModels}). A value of $\fc=0.01$ for ${f_{\!\mathrm{\mu}}}=0$ is also
supported by calibrations of the lithium, beryllium and boron surface
abundance for the sun by {\cite{Fli93}}.
\section{Numerical solution}
\lSect{NumSolve}
Two different numerical codes were used here to follow the stellar
evolution. We now briefly describe each.
\subsection{STERN}
\lSect{STERN}
The STERN code is a pseudo-Lagrangian, implicit hydrodynamic code
{\Cite{lan88}}, based on the ``G\"ottinger stellar evolution code''.
For numerical solution, relative mass coordinate $q:=m/M$ is used
instead of the the mass coordinate $m$, which allows to reserve the
distribution of computational grid in the presence of mass loss.
The equation of state includes radiation, ionization, relativistic
electron degeneracy, and electron-positron pairs. Ions are
treated as a Boltzmann gas {\Cite{EL86}}.
The chemical evolution due to thermonuclear burning is traced by $35$
isotopes: {\I{}n}, {\I{1,2}H}, {\I{3,4}{He}}, {\I{6,7}{Li}},
{\I{7,9}{Be}}, {\I{8,10,11}B},{\I{11,12,13}C},{\I{12,14,15}N},
{\I{16,17,18}O}, {\I{19}F}, {\I{20,21,22}{Ne}}, {\I{23}{Na}},
{\I{24,25,26}{Mg}}, {\I{26,27}{Al}},{\I{28,29,30}{Si}}, and
{\I{56}{Fe}}. Except for {\I{19}F}, {\I{26}{Al}}, and {\I{56}{Fe}},
reactions between them are solved in a $32$ isotope network. These
reaction rates are also used to determine the nuclear energy
generation rate. The {\El{Ne}}/{\El{Na}} and {\El{Mg}}/{\El{Al}}
hydrogen-burning cycles are solved separately using a $13$ isotope
network including {\I1H}, {\I{18}O}, {\I{19}F}, {\I{20,21,22}{Ne}},
{\I{23}{Na}}, {\I{24,25,26}{Mg}}, {\I{26,27}{Al}}, {\I{28}{Si}}, and
{\I{16}{O }} {\Cite{Braun97}}. The neutrino losses are determined
according to {\cite{MKI85}}.
The reaction networks are solved separately for each zone between the
individual stellar structure integration time-steps. This allows for
subcycling of the reaction network with fine time-steps wherever
needed.
\subsection{KEPLER}
\lSect{KEPLER}
In the KEPLER code {\Cite{WZW78,WWF84,WW88}} the equation of state
includes a crude treatment of Coulomb corrections, beyond what is used
in STERN {\Cite{WZW78}}. A $19$-isotope network is employed through
oxygen burning, including the elements {\I1H}, {\I3{He}}, {\I4{He}},
{\I{12}C}, {\I{14}N}, {\I{16}O}, {\I{20}{Ne}}, {\I{24}{Mg}},
{\I{28}{Si}}, {\I{32}S}, {\I{36}{Ar}}, {\I{40}{Ca}}, {\I{44}{Ti}},
{\I{48}{Cr}}, {\I{52}{Fe}}, {\I{54}{Fe}}, {\I{56}{Ni}} and neutrons
and protons from photodisintegration. Silicon burning is followed
using a quasi-equilibrium network of $137$ isotopes, in which
subgroups of elements are treated in nuclear statistical equilibrium
while reactions between these subgroups are considered explicitly.
Beyond silicon burning full nuclear statistical equilibrium is
assumed. A more detailed description of the reaction networks in
KEPLER can be found in {\cite{WZW78}}. However, an improvement of the
treatment of hydrogen burning has been implemented (\App{KepImpr}).
For the present work, angular momentum has been added to KEPLER as a
new local variable, and rotationally induced mixing processes
incorporated according to {\Sect{RotMix}}. However, because changes
to the structural model calculations on KEPLER would be difficult, the
modifications to the momentum balance and the energy transport
({\Sect{ModRot}}) applied in STERN are not included into KEPLER. The
same opacities {\Cite{IR96}} used in STERN are also included in KEPLER
(an update to previous versions of the code), which allows for more
consistency between the two calculations. For temperatures above
$10^9\;{\mathrm{K}}$ the opacities used in KEPLER are still chiefly due to
electron scattering with corrections due to relativity and degeneracy
{\Cite{WZW78}}.
As outer boundary conditions a finite (or zero) boundary pressure is
often utilized in KEPLER. The radius of the photosphere is determined
as the location where an optical depth of $2/3$ is reached. This
treatment of the outer boundary condition, but also the mass loss, is
less accurate than that implemented in STERN {\Cite{Heg98}}. For this
reason, the stellar evolution from the pre-main sequence until a
central temperature of $10^9\,{\mathrm{K}}$, i.e., before central neon ignition,
is followed by STERN, and the rest of the evolution until core
collapse by KEPLER. At this stage of evolution, the total mass lost
in its remaining lifetime ($\simle100\,\yr$) prior to core collapse is
negligible. The stellar envelope, and therefore the outer appearance
of the star, hardly changes. However, stellar models followed form
the pre-main sequence using KEPLER give results similar to those
obtained by STERN.
\section{Initial models}
\lSect{InitialCond}
The initial model for the calculations presented in this work is that
of a fully convective, rigidly rotating (following our assumption that
convection does lead to rigid rotation) pre-main sequence star. In the
Hertzsprung-Russell (HR) diagram such stars are located on their
Hayashi line. These models are constructed from the Lane-Emden
equation {\Ce{KW91}} with a polytropic index of $n=3/2$. Typically,
initial stellar radii around $1\,000\,{\mathrm{R}_{\odot}}$ are used. This kind of
initial condition is for computational convenience only and is not
intended to reproduce the true pre-main sequence evolution
{\Csa{BM94,BM96}}.
The influence of rotation on the stellar structure is negligible in
the initial models, but it becomes more important when the stars
contract towards central hydrogen ignition. On the zero-age main
sequence (ZAMS) close-to rigid rotation establishes throughout the
star, mainly through the action of Eddington-Sweet circulation
(\Sect{ES}) and the Gold\-reich\--Schu\-bert\--Fricke instability
(\Sect{GSF}). These processes are sufficiently efficient in the early
stellar evolution that rigid rotation is established virtually
independent of the initial angular momentum distribution assumed.
Almost no angular momentum is lost before the star reaches the main
sequence.
\pTab{InitAbuSTERN}
\pTab{InitAbuKEPLER}
All models in this work use an approximately solar initial chemical
composition with a mass fraction of all elements heavier than helium
(``metals'') of $Z=0.02$. The mass fractions of hydrogen and helium
are set to $X=0.7$ and $Y=1-X-Z=0.28$, respectively. In STERN
(\Sect{STERN}), the abundance ratios of the isotopes within each of
these groups are chosen to have the solar system meteoritic abundance
ratios according to {\cite{GN93}} (see {\Tab{InitAbuSTERN}}).
Calculations performed with the KEPLER code (\Sect{KEPLER}) start
on the pre-main sequence with a relative distribution of the metals
according to {\cite{AG89}} as given in {\Tab{InitAbuKEPLER}}.
For the main set of models in this work the initial angular momentum
is determined such that the stars reach a rotational velocity of
$\sim200\,{\km\,\Sec^{-1}}$ on the ZAMS. This is a typical observed value for
stars in the mass range $8\,{\mathrm{M}_{\odot}}\ldots25\,{\mathrm{M}_{\odot}}$
{\Cite{Sle70,Fuk82,Hal96,Pen96,how97}}. It corresponds to $\sim
35\,\%$ of their ``critical'' rotation speed (\Sect{Omega}). Also
models with different initial rotation rates are computed, in order to
investigate the influence of this parameter on the evolution of
massive stars (see {\Tab{InitModels}}).
\section{Central hydrogen burning}
\lSect{cHburn}
\subsection[Chemical mixing: the example of $20\,{\mathrm{M}_{\odot}}$ stars]
{Chemical mixing: the example of $\mathbf{20\,{\mathbf{M_{\odot}}}}$ stars}
\lSect{IntHydroMix}
\pFig{m-X-20AB}
In {\Fig{m-X-20AB}} the internal profiles of the most abundant
isotopes in a non-rotating star and two rotating $20\,{\mathrm{M}_{\odot}}$ models
are compared at core hydrogen exhaustion. Convection causes flat
profiles in the innermost few solar masses. Small convective and/or
semiconvective regions (similar to {\ModA{D15}} in {\Fig{D15cnv}})
cause steps in the profile above the convective core.
In the non-rotating case no mixing occurs in the envelope. In
contrast, the rotating models mix thermonuclear processed matter into
the envelope. If no inhibition of rotationally induced
instabilities by $\mu$-gradients is assumed an extended gradient in
helium (along with other species) reaches from the upper edge of the
convective core up to the surface ({\ModA{E20}} in
{\FIG{A}{m-X-20AB}}). Due to the increase of the mean molecular
weight in the whole envelope, as a consequence of helium enrichment,
the mass of the hydrogen-depleted core of {\ModA{E20}} is about
$1.5\,{\mathrm{M}_{\odot}}$ larger than in the non-rotating case.
The dominant rotationally induced mixing process during central
hydrogen burning is Eddington-Sweet circulation. It is fast
enough to keep the whole star close to rigid rotation
(\Sect{MSjtrans}), and thus renders shear instabilities unimportant.
The GSF instability remains one to two orders of magnitude less
efficient than the Eddington-Sweet circulation. The $\mu$-gradients
above the convective core
in {\ModA{E20}} (see also {\Fig{E15cnv}}) are strong enough to
suppress the occurrence of extended semiconvective structures. The
secular shear instability occurs only in a small layer close to the
surface, and never contributes significantly to the mixing.
\lSect{MixHmu}
If $\mu$-gradients {\emph{are}} taken into account for the
rotationally induced instabilities ({\ModA{E20B}};
{\FigBB{B}{m-X-20AB}{}{E15Bcnv}}), the $\mu$-gradient which forms at
the upper edge of the convective core is not smoothed out fast enough,
but instead almost completely chokes off any mixing between core and
envelope quite early during core hydrogen burning. Therefore, below
$m=10\,{\mathrm{M}_{\odot}}$ the composition of {\ModA{E20B}} remains quite similar
to that of {\ModA{D20}}. The higher concentration of carbon in
{\ModA{E20B}}, however, shows the occurrence of {\emph{some}} mixing
early on.
Above the ``barrier'' due to the $\mu$-gradient ($\mu$-barrier) mixing
is efficient (see the small slope of the composition profiles in the
envelope of {\ModA{E20B}}; {\FIG{B}{m-X-20AB}}), and stronger than for
{\ModA{E20}}, since the efficiency for compositional mixing is assumed
to be $\fc=1/30$ in {\ModA{E20B}} instead of $\fc=1/100$ for
{\ModA{E20}}.
The relative contributions of the different rotationally induced
mixing processes above the $\mu$-barrier are similar in
{\ModB{E20}{E20B}}, except that close to the $\mu$-barrier the GSF
instability becomes important in {\ModA{E20B}}. Within the
$\mu$-barrier, almost all rotationally induced mixing is suppressed
and the mixing is dominated by semiconvection. The secular shear
instability is
inhibited by the $\mu$-gradient.
Strong angular velocity gradients at the boundaries of convective layers
cause, in principle, layers where the shear flow can overcome
the stabilizing effect of the $\mu$-gradients. However, they are
too thin to be resolved in the present calculations.
\subsection{Transport of angular momentum}
\lSect{MSjtrans}
\pFig{m-wjJ53-E15AB}
Similar to chemical mixing, the transport of angular momentum depends
strongly on the inhibition of rotationally induced mixing by
$\mu$-gradients. {\Fig{m-wjJ53-E15AB}} compares the internal angular
velocity profile of two $15\,{\mathrm{M}_{\odot}}$ stars ({\ModB{E15}{E15B}}) which
were computed with different values of ${f_{\!\mathrm{\mu}}}$ (\Sect{calib}).
In {\ModA{E15}}, the difference between surface and core angular
velocity remains less than $30\,\%$ during core H-burning. The
over-all decrease of the rotation rate by a factor of $\sim3$ is
caused by two effects: mass loss from the surface, which carries away
$\sim40\,\%$ of the initial angular momentum, and the expansion of the
stellar envelope, which increases the total moment of inertia by a
factor of $\sim2$. At the same time, the stellar core contracts. The
persistence of almost rigid rotation during core hydrogen burning
implies transport of angular momentum from the core to the envelope.
This is confirmed by {\FIG{C}{m-wjJ53-E15AB}} which shows a decrease
of the core specific angular momentum with time (see also
{\FIG{E}{m-wjJ53-E15AB}}). Because of its small radial extent, the
core contains only a small fraction of the total angular momentum of
the star (\FigBb{E}{m-wjJ53-E15AB}{F}). For higher mass loss rates,
i.e., for more massive stars, the spin-down (decrease of $\omega$) is
dominated by the mass loss, while at lower mass it is dominated by the
expansion of the envelope.
{\FIG{B}{m-wjJ53-E15AB}} shows that the inhibition of rotational
mixing leads to differential rotation during core hydrogen burning.
The ratio of the core to envelope angular velocity in {\ModA{E15B}}
becomes $\sim4$ at core hydrogen exhaustion. The envelope rotates
slightly faster than in {\ModA{E15}} since the star loses only
$20\,\%$ of the initial total angular momentum, i.e., about half as
much as {\ModA{E15}}. This is due to the lower luminosity of
{\ModA{E15B}} during core hydrogen burning --- due to less efficient
chemical mixing (\Sect{IntHydroMix}) --- and consequently about
$60\,\%$ less mass loss than in {\ModA{E15}}.
{\FigBB{D}{m-wjJ53-E15AB}{F}{m-wjJ53-E15AB}} show that in
{\ModA{E15B}} the core angular momentum is constant throughout core
hydrogen burning.
\lSect{J53def}
{\FigBb{E}{m-wjJ53-E15AB}{F}} compare the angular momentum
distribution of {\ModA{E15}} and {\ModA{E15B}} at various evolutionary
stages using the variable ${J\Brak{m}}/m^{5/3}$, with $J\Brak{m}:=\int_0^m
j\Brak{m'}{\,\D\!\!\;} m'$. Since for a rigidly rotating body of constant
density, $\rho_0$, the angular momentum ${J\Brak{m}}$ enclosed by the mass
coordinate $m$ is
\begin{equation}
J\Brak{m} =\frac{3\omega k}{5}\Brak{\frac{3}{4\pi\rho_0}}^{\!2/3}
m^{5/3}
\propto m^{5/3}
\;
\lEq{J53def}
\end{equation}
the curves in {\FigBb{E}{m-wjJ53-E15AB}{F}} are more or less flat.
The evolution of ${J\Brak{m}}$ illustrates the transport of angular momentum
throughout stellar evolution. ${J\Brak{m}}$ drops when angular momentum is
transported through the mass shell $m$. If no transport angular
momentum through the mass shell $m$ occurs, ${J\Brak{m}}$, and also
${J\Brak{m}}/m^{5/3}$, remain constant. Furthermore, following a line of
constant $J$ from one evolutionary stage to a subsequent one shows to
what mass coordinate angular momentum has been transported in the star
during the time between the two evolutionary stages. We will refer
more to {\FigBb{E}{m-wjJ53-E15AB}{F}} in the discussion of the angular
momentum transport during the later evolutionary phases.
\subsection{The influence of the initial rotation rate}
\lSect{inflji}
\pFig{m-wjJ53-GF15B}
{\Fig{m-wjJ53-GF15B}} shows the evolution of angular velocity and
specific angular momentum in {\ModB{G15B}{F15B}}, which both contain
the inhibition of rotational mixing due to $\mu$-gradients. The
latter model initially has three times more angular momentum than the
first.
While this difference of a factor of three in the rotation rate is
conserved in the envelope throughout core hydrogen burning, it becomes
much smaller in the cores. The faster rotation of {\ModA{F15B}}
sustains the transport of angular momentum out of the core for a
longer time than in {\ModA{G15B}}, where the core angular momentum is
almost completely conserved (\Fig{m-wjJ53-GF15B}). That is, the
angular momentum is less efficiently trapped in the fast rotating
{\ModA{F15B}} than in the {\ModB{G15B}{E15B}}
(\FigB{m-wjJ53-E15AB}{m-wjJ53-GF15B}). This feedback process leads to
a convergence of the core rotation rates. We note already here that
this convergence persists during the later burning stages and leads to
very similar iron core angular momenta for a wide range of initial
rotation rates (cf.~{\Sect{j:preSN}} below).
The stronger core angular momentum depletion in faster rotating models
occurs simultaneously with rotationally induced mixing across the
$\mu$-barrier: The masses of the convective cores at the end of
central hydrogen burning are $2.4\,{\mathrm{M}_{\odot}}$, $2.5\,{\mathrm{M}_{\odot}}$, $2.6\,{\mathrm{M}_{\odot}}$,
and $2.8\,{\mathrm{M}_{\odot}}$ for {\ModD{D15}{G15B}{E15B}{F15B}}, respectively.
However, {\ModA{E15}}, where the $\mu$-barrier was assumed to be
inefficient, has a core of about $3.5\,{\mathrm{M}_{\odot}}$. Thus, even for very
rapid rotation the assumption of $\mu$-barriers inhibiting rotational
mixing strongly restricts the core growth due to rotation. (see also
{\FigC{D15cnv}{E15cnv}{E15Bcnv}}).
\section{Central helium burning}
\lSect{cHeBurn}
After core hydrogen exhaustion, the models become red supergiants
(except for {\ModA{H12B}} which first burns helium as a as a blue
supergiant for some time) and their extended hydrogen-rich envelopes
become convective. The pulsational properties of these envelopes have
been discussed by {\cite{heg97}} and the evolution of the surface
rotation rates, especially during blue loops, by {\cite{HL98}}. In
the following, we investigate the evolution of the cores, using
the $15\,{\mathrm{M}_{\odot}}$ models as example.
\pTab{tauES}
The importance of rotation in the post main sequence evolution can
be estimated from {\Tab{tauES}}, which compares the
Eddington-Sweet time-scale
\begin{equation}
{\tau_{\mathrm{ES}}}\sim{\tau_{\mathrm{KH}}}\Brak{\frac{\wk}{\omega}}^{\!2}
\;,\qquad
\wk:=\sqrt{Gm/r^3}
\;.
\end{equation}
{\Cite{Zah92}} in the cores of our {\ModB{E15}{E15B}} during the
various burning stages with the respective nuclear time scales. For
the amount of differential rotation in our models, the characteristic
time-scale for mixing due to the GSF instability (\Sect{GSF}) is
comparable to the Eddington-Sweet time-scale.
The core hydrogen burning phase is the only one where mixing and
nuclear time scale are comparable. During core helium burning, the
mixing time scale is one or two orders of magnitude larger than the
nuclear time-scale, which may still allow for some effects of
rotational mixing. The later phases are too short to allow for any
rotationally induced mixing in the cores; note however, that at the
core boundaries some effects of rotational mixing may still be
possible in case of strong gradients in the angular velocity
(cf.~\Sect{hydMix} below).
An energetic limit to the amount of mixing due soley to shear
instabilities can be obtained by comparing the rotational energy of
the core with the potential energy required to lift processed matter
from the upper edge of the convective core to the hydrogen-burning
shell source {\Cite{Heg98}}. For a typical value of $\omega/\wk=0.05$
and a difference in the mean molecular weight of fully ionized carbon
relative to helium of $\sim0.3$ (oxygen would be even heavier), an
enrichment of carbon by at most $\simle0.5\,\%$ is possible. This
assumes the carbon to be homogeneously distributed throughout the
radiative layer and that all the rotational energy of the core is used
to supply the buoyancy energy. Note that this limit does not apply to
instabilities which tap the energy flux in the star like the
Eddington-Sweet circulation.
\subsection{Chemical mixing}
\lSect{HeMix}
\pFig{m-X-DE15AB}
In a non-rotating star using the Ledoux criterion for convection
(\Sect{Semiconv}) prevents the growth of the convective helium core
that would occur if the Schwarzschild criterion were assumed.
Instead, several convective regions, separated by semiconvective
layers, form above the convective core
({\FigBB{C}{m-X-DE15AB}{}{D15cnv}}). In the rotating models with
${f_{\!\mathrm{\mu}}}=0.05$ (e.g., {\ModA{E15B}} in
{\FigBB{B}{m-X-DE15AB}{}{E15Bcnv}}) the shear across the
semiconvective layers is not strong enough to overcome the stabilizing
$\mu$-gradient, even for the fast rotating {\ModA{F15B}}.
If rotationally induced mixing is assumed to be insensitive to
$\mu$-gradients (i.e., ${f_{\!\mathrm{\mu}}}=0$; {\ModA{E15}} in
{\FigBB{A}{m-X-DE15AB}{}{E15cnv}}) the dynamical shear instability
operates in the semiconvective regions and dissolves them, similar to
the case of Schwarzschild convection. In this case, the rotational
mixing leads to considerably more massive helium cores. The
resulting higher burning temperatures in the cores lead to lower
central carbon-to-oxygen ratios at core helium exhaustion.
An interesting issue is the mixing (and angular momentum transport) in
the radiative helium layer between the convective core and the
hydrogen-burning shell. If the products of helium burning could be
mixed upward into the hydrogen-burning shell, {\emph{primary}}
production of {\I{14}{N}} could occur. If hydrogen were transported
down into the helium-burning center, a much stronger than normal
s-process could result and build up more heavy or neutron-rich
elements. On the other hand, strong instabilities in this region
could also lead to a significant slowing-down of the core.
The dominant mixing process present in this layer is Eddington-Sweet
circulation, with some contribution from the GSF instability. During
the early stages of core helium burning of models with ${f_{\!\mathrm{\mu}}}=0$ (e.g.,
{\ModA{E15}}), the secular shear instability dominates slightly over
the Eddington-Sweet circulation at the upper edge of the helium core.
Towards central helium exhaustion, the mixing is dominated by the GSF
instability. In the case of ${f_{\!\mathrm{\mu}}}=0.05$, the secular shear
instability is suppressed by $\mu$-gradients.
{\FIG{B}{m-X-DE15AB}} illustrates that some mixing occurs during core
helium burning: A gradient in {\I{12}C} and {\I{16}O} extends from the
convective core up to the edge of the helium core. In this model, the
increase in {\I{12}C} or {\I{16}O} is not sufficient to result in any
significant primary nitrogen production in the hydrogen burning
shell. Even though this effect is not notably more pronounced in the
initially faster rotating {\ModA{F15B}} --- due to the convergence of
the core rotation rates; cf.~\Sect{inflji} --- or for the different
initial masses investigated here, such a primary nitrogen production
appears possible in more favorable conditions, e.g., for higher
mixing efficiencies or at lower metallicity.
In {\ModA{E15}} (\FIG{A}{m-X-DE15AB}) the rotation of the helium core
is slower, and the {\I{12}C} and {\I{16}O} gradients are much steeper,
leveling off to the CNO equilibrium values a few tenths of a solar
mass above the convective core. In the non-rotating {\ModA{D15}}
(\FIG{C}{m-X-DE15AB}), no enrichment of {\I{12}C} and {\I{16}O}
appears at all above the outermost semiconvective layer of the
convective core.
Even though the strong entropy gradient at the location of the
hydrogen-burning shell suppresses rotational mixing between the helium
core and the hydrogen burning shell, some mixing occurs due to the
large angular velocity gradient. This can be seen in {\Fig{E15Bcnv}}:
The tail of the energy generation rate at the lower bound of the
hydrogen-burning shell source in {\ModA{E15B}} penetrates into the
helium core, i.e., some hydrogen is mixed downward. Since the protons
burn quite fast as they are mixed deeper inside the helium core, they
cannot reach the central convective region. However, some protons may
survive and get mixed into the convective helium shell later on
(\Sect{hydMix}). In {\ModA{E15}} (\Fig{E15cnv}), where the core is
rotating slower, and also in the non-rotating {\ModA{D15}}
(\Fig{D15cnv}), this feature is not found.
\pFig{E15Bcnv}
\subsection{Transport of angular momentum}
\lSect{HeAngTrans}
After core hydrogen exhaustion, the stars undergo a phase of major
restructuring as the core contracts and the envelope expands. This
leads to a spin-up of the core
({\FigDdd{A}{m-wjJ53-E15AB}{B}{A}{m-wjJ53-GF15B}{B}}) and a spin-down
of the envelope. At the same time, the convective envelope grows in
mass and its bottom approaches the helium core. A steep rise in the
specific angular momentum occurs at the bottom of the rigidly rotating
envelope that persists throughout core helium burning and beyond
({\FigDdd{C}{m-wjJ53-E15AB}{D}{C}{m-wjJ53-GF15B}{D}}). The entire
helium core stays close to rigid rotation during central helium
burning ({\FigDdd{A}{m-wjJ53-E15AB}{B}{A}{m-wjJ53-GF15B}{B}}).
Up to core helium exhaustion, the specific angular momentum of the
helium core drops appreciably with time
(\FigB{m-wjJ53-E15AB}{m-wjJ53-GF15B}). Three processes contribute to
this effect. First, angular momentum is removed from the core during
the star's restructuring phase between core hydrogen depletion and
helium ignition. Second, the core grows in mass due to hydrogen shell
burning and engulfs regions of lower specific angular momentum
(\FigBb{C}{m-wjJ53-E15AB}{D}). The reasons for the low specific
angular momentum above the core are secular shear instabilities, the
{\emph{first dredge-up}}, and short-lived convective regions which
temporarily extend down to mass coordinates smaller than the final
helium core mass. The regions of outwards decreasing specific angular
momentum are not Solberg-H{\o}iland unstable due to strong stabilizing
entropy and composition gradients. Third, some angular momentum is
transported from the helium core into the envelope through the
hydrogen-burning shell. The models of the ``\Mod{B}'' series lose
less angular momentum during the restructuring phase because of the
inhibiting effect of the $\mu$-gradients, but more during central
helium burning, due to their considerably faster rotation.
The relative loss of angular momentum in {\ModC{G15B}{E15B}{F15B}}
during helium burning increases with the initial amount of angular
momentum left at the end of central hydrogen burning. Consequently
all three models end up with very similar core angular momenta and
rotation rates (\Fig{m-wjJ53-GF15B}), about three times that of
{\ModA{E15}} (\Fig{m-wjJ53-E15AB}).
\section{Comparison with previous work}
\lSect{CmpOther}
In contrast to {\cite{KMT70}}, who investigated rapidly rotating
$9\,{\mathrm{M}_{\odot}}$ stars ($v\simgr400\,{\km\,\Sec^{-1}}$), our models do not become
secularly unstable at the end of central helium burning, since,
according to our assumptions, $\mu$-barriers are less efficient in
suppressing angular momentum transport compared to {\cite{KMT70}}.
{\cite{ES78}} followed the evolution of $7\,{\mathrm{M}_{\odot}}$ and $10\,{\mathrm{M}_{\odot}}$
stars with a ZAMS rotational velocity of $\sim200\,{\km\,\Sec^{-1}}$ using
essentially the same method as in the present work, except for some
improvements in the input physics of the individual processes applied
here {\CITE{{\Sect{MixInst}} and}{ES78}}. They used the Schwarzschild
criterion for convection, however, and did not include mass loss. In
their work, the $\mu$-barrier above the convective hydrogen-burning
core suppressed mixing and transport of angular momentum almost
completely. Therefore, their stellar cores lose very little angular
momentum during central hydrogen burning. Although we use ${f_{\!\mathrm{\mu}}}<1$
and the inhibiting effect of the $\mu$-gradients is smaller, a similar
$\mu$-barrier forms during central hydrogen burning. However, we
obtain some mixing between the core and the envelope early during core
hydrogen burning, some angular momentum loss from the core to the
envelope, and in most cases some enrichment of the surface with
H-burning products. In an earlier work, {\cite{ES76}} disregarded
rotationally induced angular momentum transport, but imposed various
rotation laws. In this case an even more extreme result was obtained:
all models reached critical rotation before carbon ignition.
{\cite{ery94}} considered turbulent diffusion according to
{\cite{Zah83}} in their computation of a rotating $20\,{\mathrm{M}_{\odot}}$ stars
with a metallicity of $Z=0.008$. They found a surprisingly large
surface {\I{14}N} enrichment at the end of core hydrogen burning of
more than $2\,\%$ by mass. Since the CNO cycle conserves the total
mass of the CNO isotopes, this result appears implausible and cannot
be reproduced in the present work.
{\cite{MM97}} used a prescription for the Eddington-Sweet circulation
according to {\cite{Zah92}}, and a modified Richardson criterion to
account for thermal effects {\Cite{Mae95,MM96}}. They computed the
hydrogen-burning evolution of stars from $9\,{\mathrm{M}_{\odot}}$ to $60\,{\mathrm{M}_{\odot}}$ and
found a strong inhibiting effect of the $\mu$-gradients on the
rotational mixing, which resulted in stronger differential rotation at
core hydrogen exhaustion than found in our models of the ``\Mod{B}''
series. Their models did not show any surface enrichment of helium.
These models were superseded by those of {\cite{Meyn97}}, who computed
the main sequence evolution of $20\,{\mathrm{M}_{\odot}}$ and $40\,{\mathrm{M}_{\odot}}$ stars,
using improved physics of rotationally induced mixing as discussed by
{\cite{MM97}}, and {\cite{Mae97A}}. The inhibiting effect of
$\mu$-gradients on shear mixing and Eddington-Sweet circulation was
strongly reduced in the new formulation. {\citeauthor{Meyn97}}'s
rotating $20\,{\mathrm{M}_{\odot}}$ model showed a larger envelope helium enrichment
than a comparable models of {\ModA{E20}}, and a similar mass of the
hydrogen-depleted core. The physics used in {\cite{Meyn97}} has been
revised again by {\cite{MZ98}} for a more consistent treatment of
$\mu$-gradients. Models with this prescription are not yet available.
{\cite{tal97}} followed the main sequence evolution of $9\,{\mathrm{M}_{\odot}}$
stars with ZAMS rotation rates of $100\,{\km\,\Sec^{-1}}$ and $300\,{\km\,\Sec^{-1}}$ until
end of central hydrogen burning, using the prescription for the
Eddington-Sweet circulation by {\cite{Zah92}}. The helium enrichment
in the envelope showed a smooth profile, similar to our
{\ModB{E08}{E10}}. At core hydrogen exhaustion, their models showed
steep composition gradients close to the stellar surface. This may have
resulted from the low mass loss assumed in their calculation in
combination with inefficient mixing close to the surface. With a
slightly larger mass loss rate, their rapidly rotating model would
have a much stronger surface enrichment. Due to the downward
advection of angular momentum by meridional circulation in the theory
of {\cite{Zah92}}, {\cite{tal97}} found a somewhat stronger envelope
differential rotation compared to our {\ModB{E08}{E10}}, but a
comparable one to, e.g., {\ModA{E12B}}. We conclude that this
downward advection is not a strong effect, which may justify its
neglect in the present work.
Summarizing, the prescription for rotationally induced mixing used in
the model of {\cite{KMT70}} corresponds roughly to ${f_{\!\mathrm{\mu}}}=\fc=\infty$
in terms of the present formulation. {\cite{ES78}} used about
${f_{\!\mathrm{\mu}}}=\fc=1$. Neither work obtained any surface enrichment during
core hydrogen burning due to the strong inhibiting effect of
$\mu$-gradients {\Cite{MM97}}. The recent picture of interacting
Eddington-Sweet circulation, anisotropic turbulence and shear
instabilities {\Cite{CZ92,MZ98}} has been continuously improved in the
last years {\Ce{USS96,Mae97A,TZ97,tal97}}. The most recent work in
this series, {\cite{MZ98}}, includes an improved treatment of
compositional gradients, but this type of description for rotationally
induced mixing is complex, computationally expensive, and has not yet
been successfully tested for post-hydrogen burning stars. However,
the results obtained in earlier work {\Cite{tal97,Meyn97}} for massive
main sequence stars are not significantly different from those of the
present work.
\section{Late evolution until core collapse}
\lSect{LateEv}
After core helium exhaustion, the carbon-oxygen core ({CO} core)
contracts and subsequently phases of carbon, neon, oxygen, and silicon
central convective and shell burning follow inside this core.
{\Tab{FinalModels}} gives some key parameters of the final models: the
final mass of the star, the masses of the helium, {CO}, and iron
cores, and the angular momenta contained in theses cores. For the
iron core additionally the average specific angular momentum is given.
{\ModF{D10}{D12}{E08}{G12B}{E12B}{F12B}} develop degenerate
neon-oxygen cores and central neon burning starts off-center. Due to
the computational difficulties (and expense) these models were not
followed until core collapse, but stopped during neon or oxygen shell
burning or even before neon ignition. In {\ModA{D10}} even carbon
burning ignites off-center. {\ModB{D10}{E08}} experience a dredge-up
of almost the entire helium shell by the convective envelope.
Therefore the final helium core is small and has little angular
momentum. The remaining helium shell above the {CO} core at the
point where the calculation is ended is only a few hundreths of a
solar mass. The masses of the helium cores before the dredge-up are
$2.25$ and $2.1\,{\mathrm{M}_{\odot}}$, respectively. {\ModA{E25}} loses its
hydrogen-rich envelope during central helium burning and becomes a
Wolf-Rayet star. Strong Wolf-Rayet mass loss sets in and further
decreases the mass of the star. It ends up with only $5.45\,{\mathrm{M}_{\odot}}$ at
the time of core collapse and very little angular momentum.
Note that in {\Tab{FinalModels}} the masses of some of the helium
cores in the non-rotating models are apparently larger than those of
the slowly rotating models of the ``\Mod{B}'' series. This is an
artifact due to the criterion used to measure the mass of the helium
core. We define the helium core by the mass coordinate at which the
hydrogen mass fraction drops below $\Ep{-3}$. In the rotating models
of the ``\Mod{B}'' series, the hydrogen gradient at the top of the
helium core is significantly shallower. If instead a hydrogen mass
fraction of $\Ep{-2}$ is chosen, the helium core masses are similar or
even larger for the rotating models. Note that the size of the {CO}
core is defined in a similar way: by the mass coordinate at which the
helium mass fraction drops below $\Ep{-3}$.
\subsection{Chemical mixing in the helium shell}
\lSect{hydMix}
After core helium exhaustion, the {CO} core contracts and the burning
of helium continues in a shell. At the same time, the outer layers of
the helium-rich shell cool down and the hydrogen shell source goes
out. Since this implies a reduced entropy barrier, rotationally
induced mixing through the hydrogen-helium interface can now operate
more efficiently. The protons which are mixed downward into the
helium shell do not burn immediately. When later the helium-burning
driven convective shell extends upwards, it dredges these protons down
into the hot, helium-burning layers (see {\Fig{E15Bcnv}}). This
mechanism can open new channels of nucleosynthesis. This will be
investigated in more detail in forthcoming papers {\CS{ for a first
report}{lan99:CosV}}.
\subsection{Chemical mixing inside the {CO} core}
During the final remaining stellar burning phases, rotational mixing
inside the {CO} core is unimportant. The strongest instabilities are
again the Eddington-Sweet circulation and the GSF instabilities, at
about same order of magnitude. From {\Tab{tauES}} it can be seen that
their time scale is too long in order to be significant.
Also, the mixing of {\emph{traces}} of material into regions
of neighboring burning phase is not expected to introduce
qualitatively new nucleosynthesis channels, since all
abundant nuclear species in one burning phase are anyway present
in the neighboring one (e.g., mixing traces of oxygen into neon
burning is not exciting).
In the fast rotating cores of the models of the ``{\Mod{B}}'' series,
secular shear instabilities arise above several of the central and
shell convection zones for a limited time, but they do not become
efficient enough to cause any noticeable mixing.
\subsection{Transport of angular momentum}
\lSect{lateJtrans}
As for the chemical mixing, rotational mixing cannot effectively
remove angular momentum from the core during the late burning stages.
In particular, transport is too inefficient to keep the {CO} core in
rigid rotation. Strong differential rotation occurs
({\FigDdd{A}{m-wjJ53-E15AB}{B}{A}{m-wjJ53-GF15B}{B}}). At this time,
the only instability capable of enforcing rigid rotation is
convection. Since the radii of the lower boundaries of the major
convection zones of carbon, oxygen, and silicon burning are much
smaller than that of their upper edges, large differences in the
specific angular momentum exist between the bottom and the top of the
convection zone. Thus, angular momentum is mainly carried outwards.
The typical signature of such a convection zone is a steep drop of the
specific angular momentum at its bottom, accompanied by a large
increase at its top (\FigC{m-wjJ53-E15AB}{m-wjJ53-GF15B}{jj15B}).
Convection zones that subsequently overlap can transport angular
momentum efficiently over scales larger than their individual extent.
This is most efficient when the lower boundary of a convective shell
overlaps with the upper boundary of a preceding convection zone. For
the models investigated in this work, such an overlap occurs rather
infrequently ({\App{convDiag}}). Subsequent shells, which are
driven by nuclear burning, tend to form their lower boundary at the
upper edge of a preceding shell, where the fuel for their burning is
not yet depleted. The most prominent example of this is the
sequence of carbon-burning shells (e.g., {\Fig{E15cnv}}).
Exceptions occur only for some of the late carbon burning shells, and
for the oxygen burning shells in {\ModB{G15B}{E15B}}.
Convective angular momentum transport does not operate across the
boundary of the {CO} core. These cores retain their angular momentum
after core helium exhaustion. Some redistribution, mainly due to
convection, occurs inside the cores. For example,
{\FIG{E}{m-wjJ53-E15AB}} shows that after core helium exhaustion in
{\ModA{E15}}) no angular momentum is transported through the shells at
$m\approx 3.4\,{\mathrm{M}_{\odot}}$ and $m\approx5.1\,{\mathrm{M}_{\odot}}$, i.e., the boundaries
of the helium and the {CO} core, respectively.
In models with more rapidly rotating cores,
({\ModC{G15B}{E15B}{F15B}};
{\FigCCc{F}{m-wjJ53-E15AB}{E}{m-wjJ53-GF15B}{F}}), the helium core
does lose some angular momentum, even though its upper boundary (at
$m\approx3.7\,{\mathrm{M}_{\odot}}$) remains a significant barrier for angular
momentum transport as indicated by the spike in
{\FIG{F}{m-wjJ53-E15AB}}. The loss of angular momentum from the
helium core is correlated with the mixing of hydrogen into the helium
shell described in {\Sect{hydMix}}. No significant angular momentum
was transported across the boundary of the {CO} core in any of the
models ({\FigDdd{E}{m-wjJ53-E15AB}{F}{E}{m-wjJ53-GF15B}{F}}).
\subsection{Stability to triaxial deformations}
\pFig{m-phi}
As described in {\Sect{ModRot}}, the approximations employed in this
work are limited to slow rotation in the sense that no triaxial
deformation appear. In the KEPLER code, the influence of the
centrifugal forces on the structure is completely neglected. However,
when models from calculations with STERN, where centrifugal forces
{\emph{are}} included, are continued by KEPLER at a central
temperature of $\Ep9\,{\mathrm{K}}$, the evolution usually proceeds smoothly,
i.e., these forces {\emph{are not}} important at this late stage of
evolution. On the ZAMS, the rotational energy ${E_{\mathrm{rot}}}$ of the star is
negligible in comparison to its gravitational binding energy ${E_{\mathrm{pot}}}$
for all models, even for those which are close to critical rotation at
their surface (${E_{\mathrm{rot}}}/\abs{{E_{\mathrm{pot}}}}\ll1\,\%$).
However, in the course of their evolution the stellar models contract
and --- as outlined in {\Sect{lateJtrans}} above --- the transport of
angular momentum out of the core is inhibited or slow, with the
consequence of rapidly rotating cores
({\FigDdd{A}{m-wjJ53-E15AB}{B}{A}{m-wjJ53-GF15B}{B}}, and
{\Tab{jev}}). For local angular momentum conservation in a shell with
given specific angular momentum $j$, the ratio of angular velocity to
Keplerian angular velocity scales as
\begin{equation}
\frac{\omega}{\wk}=\frac{j}{k\sqrt{Gmr}}\sim r^{-1/2}
\;,\qquad
k\approx2/3
\lEq{awawk}
\;.
\end{equation}
The ratio of the specific rotational energy to the gravitational
potential is then given by
\begin{equation}
\frac{{e_{\mathrm{rot}}}}{{\phi_{\mathrm{grav}}}}=\frac{1}{2}\Brak{\frac{\omega}{\wk}}^{\!2}
\;.
\end{equation}
This ratio is displayed for several $15\,{\mathrm{M}_{\odot}}$ pre-collapse models
in {\Fig{m-phi}}.
A uniformly rotating, self-gravitating, incompressible, and inviscid
fluid (McLaurin spheroid) becomes secularly unstable to triaxial
deformations when the ratio of rotational to gravitational potential
energy
\begin{equation}
\frac{{E_{\mathrm{rot}}}}{{E_{\mathrm{pot}}}}=\frac{1}{2}\int_0^m\omega^2(m'){\,\D\!\!\;} m' \left/
\int_0^m\wk^2(m'){\,\D\!\!\;} m'\right.
\end{equation}
exceeds $\sim0.1375$ {\Ce{OB73,Tas78}}. If this ratio exceeds
$\sim0.26$, the object becomes dynamically unstable to non-axisymmetric
instabilities and fission may occur {\Cite{OT69,OB73}}.
The stars simulated in the present work are well below these limits
even at the pre-collapse stages (${E_{\mathrm{rot}}}/{E_{\mathrm{pot}}}\simle30\,\%$ of the
critical value) and therefore no triaxial instabilities arise.
\section{Angular momentum prior to core collapse}
\lSect{j:preSN}
\pFig{jj15B}
\pFig{m-J53-presn-GEF}
Our model sequences are terminated at the onset of core collapse, defined by
the infall velocity inside the iron core exceeding $9\E{2}\,{\km\,\Sec^{-1}}$. At
this stage of evolution the investigated stars typically have central
densities of $\simle1\E{10}\,\gccm$.
From the previous discussions it is clear that the distribution of
angular momentum in the star at onset of core collapse strongly
reflects its recent convective structure.
{\Fig{jj15B}} shows the distribution of the specific angular momentum
at the pre-collapse stage of $15\,{\mathrm{M}_{\odot}}$ stars with different initial
rotation rates {\ModCx{G15B}{E15B}{F15B}}.
These three models show a very similar final angular momentum distribution
(cf. also {\FIG{A}{m-J53-presn-GEF}}), due to a similarity entire
chemical structure. The reason for this is the convergence
of the core rotation rates, i.e. their independence from the initial
rotation rates, already during hydrogen and helium burning, as outlined
in {\Sect{inflji}}.
In contrast, {\ModA{E15}} has much less
angular momentum left in the core (see also {\Tab{FinalModels}}). It
grows a larger helium and {CO} core due to the lack of sensitivity to
$\mu$-gradients.
\pTab{FinalModels}
The total angular momentum in the final models is dominated by that of
the envelope (\Tab{FinalModels}). {\ModC{G15B}{E15B}{F15B}} show that
for initially faster rotation, a slightly larger helium core
results (\Tab{FinalModels}) and therefore the stars become more
luminous. This in turn causes more mass and angular momentum
loss, which can, for the rapid rotators or for more massive stars,
decreases the total angular momentum by a larger factor
({\FigBB{B}{m-J53-presn-GEF}{B}{m-J53-presn}}).
\pFig{m-J53-presn}
\pFig{m-J53-presn4}
At hydrogen ignition, the total and even the mean specific angular
momentum of models with a given surface rotational velocity are larger
for larger initial masses (\Tab{FinalModels}). On the contrary, the
final total angular momentum decreases for larger initial mass
({\Tab{FinalModels}} and {\Fig{m-J53-presn}}). This trend is only
interrupted between $12$ and $15\,{\mathrm{M}_{\odot}}$ since our models with
initial masses of $12\,{\mathrm{M}_{\odot}}$ or less undergo a blue loop during core
helium burning which leads to an additional strong angular momentum
loss {\Cite{HL98}}. As in the limit of vanishing mass loss, the angular
momentum of our models is conserved, the decrease of the total angular
momentum for higher initial masses is solely due to the increase of
the mass loss rate for larger initial masses.
The total angular momentum of the helium and {CO} cores increases with
the the initial stellar mass of our models ({\Tab{FinalModels}}).
However, this trend is much weaker for the specific angular momentum
of the helium cores, the specific angular momenta of the {CO} cores
even decreases a little with increasing initial mass. This illustrates that
angular momentum transport from the core into the envelope is stronger
for larger cores.
Finally, we find that the specific angular momenta of the iron cores
are rather insensitive to the initial mass and rotation rate
(\Tab{FinalModels} and {\Fig{m-J53-presn4}}), due to the convergence
of the core rotation rates discussed in {\Sect{inflji}}. In the
models with ${f_{\!\mathrm{\mu}}} = 0$, angular momentum transport was efficient and
final values of ${j_{\!\isofont{Fe}}}\simeq 6\E{15}\,{\cm^2\,\Sec^{-1}}$ are found. The value for
{\ModA{E25}} is significantly lower since its {CO} core was spun down
in a Wolf-Rayet phase. The models with ${f_{\!\mathrm{\mu}}} = 0.05$ all end up with
${j_{\!\isofont{Fe}}}\simeq 1.2\E{15}\,{\cm^2\,\Sec^{-1}}$.
Note that, unless the iron core that forms after central silicon
burning is already large enough to collapse, one or more subsequent
phases of silicon shell burning occur until the critical iron core
mass is exceeded. The sizes of these shells depend on the details of
the preceding evolution. As a result, the iron core mass does not
necessarily increase monotonically with initial mass or rotation. For
example, the iron core of {\ModA{D20}} is larger than those of the
$20\,{\mathrm{M}_{\odot}}$ models of the ``\Mod{B}'' series.
\section{Implications for young pulsars and supernovae}
\lSect{youngPSR}
\pTab{jev}
{\Tab{jev}} shows, for times during the evolution, the specific
angular momentum contained in the innermost $1.7\,{\mathrm{M}_{\odot}}$ (the mass of
the iron core at core collapse) for {\ModA{E20}}. Due to the
continuous contraction of the central region of the star, it spins up
and gets closer to critical rotation {\Eqx{awawk}}. If the
pre-collapse value of the specific angular momentum is applied to a
neutron star with an assumed radius of $12\,\km$, it would rotate with
$90\,\%$ of Keplerian rotation (\Tab{jev}). {\ModA{E20}} has the
largest iron core mass of all our models (\Tab{FinalModels}), and a
lower core specific angular momentum than the models computed with
${f_{\!\mathrm{\mu}}} = 0.05$. Those models have even much more angular momentum in
the collapsing iron core than a neutron star can possibly carry
($\omega/\wk\propto jm^{-1/2}$). This much angular momentum would
certainly be important in the dynamics of core collapse, and it is
expected that significant deviations from spherical symmetry will
arise {\Cite{Ims95,aks97,ZM97,RMR98}}.
\subsection{Comparison with observed young pulsars}
\lSect{pulsCmp}
\pTab{jNS}
At $90\,\%$ of Keplerian angular velocity, the neutron star
which might form in the collapse of
the iron core of {\ModA{E20}} would have a rotation period of
$1\,{\mathrm{m}\Sec}$ (\Tab{jev}).
In {\Tab{jNS}} the periods of the four known young neutron stars
associated with supernova remnants {\Cite{mas98}} are given along with
their specific angular momentum (with the same assumption regarding
moment of inertia as above). Comparing this to the specific angular
momentum in the iron cores found in the pre-collapse models in
{\Tab{FinalModels}}, we see that the iron cores of our models have
roughly $\sim20$ to $100$ times more specific angular momentum than
found in these neutron stars. Triaxial deformations and gravitational
radiation would result --- even during the explosion. Still it might
be expected that the resulting neutron stars would spin much faster
than observed.
However, the observed ``young'' neutron stars have ages of several
hundred years. They might have spun much faster immediately
after their formation. In fact, it has been proposed recently
that rapidly
rotating hot neutron stars are spun down on a time-scale of
one year by r-mode oscillations and
accompanying emission of gravitational radiation.
These oscillations are supposed to cease at spin periods
compatible with those observed in the young neutron stars
{\Cite{LOM98,owe98}}.
Alternatively, an important angular momentum transport mechanism might
be missing in our models (see also {\Sect{OtherInst}}). {\cite{SP98}}
have assumed, without computing detailed models, that the winding up
of weak magnetic fields by differential rotation can cause enough
Maxwell stress to keep the entire star in uniform rotation until the
end of central carbon burning. This scenario approaches the problem of
the young neutron star periods from the other side: It implies initial
spin periods of $\sim100\,{\mathrm{s}}$. Since this is much larger than
observed, they employed off-center ``kicks'' during the supernova
explosion to spin them up to the observed rotation rates. This
scenario is speculative at present, since neither the evolution of
magnetic instabilities in the stellar interior nor the neutron star
kicks have been adequately investigated.
\subsection{Formation of Kerr black holes?}
\pFig{m-jLSO-E15AB}
If the large angular momenta obtained for the iron cores in this work
pose a problem for pulsars, they are very favorable for the collapsar
model for $\gamma$-ray bursts {\Cite{Woo93}}. If the cores of the
stars would collapse to a black hole, the angular momentum calculated
here would be enough to support matter in a stable disk outside
{\Cite{ST83, Nov97}}. This is indicated in {\Fig{m-jLSO-E15AB}},
where the distribution of the specific angular momentum at the
pre-collapse stage of the two $20\,{\mathrm{M}_{\odot}}$ {\ModB{E20}{E20B}} is shown.
Thin dashes and dash-dotted lines indicate the specific angular
momenta of the last stable orbit around a non-rotating and a maximum
rotating black hole with a mass equal to the mass coordinate. If the
matter in the star has more angular momentum than necessary to get
into the last stable orbit, an accretion disc must form, and
efficiently transform gravitational binding energy into heat, up to
$42.3\,\%$ of the rest mass for a maximum rotating black hole
{\Cite{ST83, Nov97}}. Note that the specific angular momentum
displayed in {\Fig{m-jLSO-E15AB}} is the latitudinal average over a
shell. Its actual value at the equator is higher than that by
$50\,\%$, while it is zero at the pole. Therefore matter might fall
in almost freely along the rotation axis, while it hits the
centrifugal barrier at the equator. In case a prompt supernova
explosion fails and a black hole forms instead of a neutron star, this
might be a mechanism for an efficient energy source for supernovae or
even a $\gamma$-ray burst {\Cite{Woo93,PWF98,MW99}}.
\section{Summary and conclusions}
\lSect{SumConcl}
We have presented the first complete numerical simulation of the
evolution of rotating stars from the ignition of nuclear burning until
the supernova stage. Emphasis has been placed on the modification of
the evolution induced by rotation. This includes an examination of
the transport processes responsible for redistributing each angular
momentum and composition and the resultant changes that occur in the
stellar structure and nucleosynthesis. The distribution of angular
momentum in the presupernova stage is of particular interest.
Two different one-dimensional hydrodynamic stellar evolution codes
were modified to include angular momentum as a new local variable.
The effects of centrifugal forces on the stellar structure were
treated in latitudinally averaged way. Rotationally induced
instabilities were included (\Sect{MixInst}): secular and dynamic
shear instabilities, the Solberg-H{\o}iland instability, the
Eddington-Sweet circulation, and the Goldreich-Schubert-Fricke
instability. The uncertain parameters of rotationally induced mixing
were calibrated using observational constraints on the surface
abundances (\Sect{calib}). Observed surface enrichments with
CNO-processed matter were reproduced for stars in the mass range from
$5\,{\mathrm{M}_{\odot}}$ to $60\,{\mathrm{M}_{\odot}}$, for typical initial stellar rotation rates.
Stellar mass loss and its dependence on the surface rotation rate were
also taken into account (\Sect{MassLoss}).
The evolution of stars of approximately solar composition in the mass
range from $10\,{\mathrm{M}_{\odot}}$ to $25\,{\mathrm{M}_{\odot}}$ was modeled up to iron core
collapse, the immediate presupernova stage.
Models that used different assumptions regarding the stabilizing
effect of gradients in the mean molecular weight on rotationally
induced instabilities were computed and compared. Observations
indicate that gradients in the mean molecular weight inhibit
rotationally induced mixing much less than in the pioneering models of
{\cite{ES78}}. This conclusion is also supported by recent
investigations of the physics of meridional circulations, shear
instabilities, and semiconvective mixing {\Cite{MZ98}}.
\subsection{Internal stellar structure}
During central hydrogen burning, the products of the burning are
mixed into the stellar envelope and new fuel is supplied to the
convectively burning stellar core by rotationally induced mixing.
Since this mixing proceeds on a time-scale comparable to the
thermonuclear time-scale of hydrogen burning, a gradient of
processed matter builds up inside the radiative envelope. The
processed matter has a higher mean molecular weight, $\mu$, than the
pristine matter of the star, and therefore a gradient of the mean
molecular weight results.
If rotationally induced mixing occurs by processes that depend
sensitively upon these gradients, they act as a barrier
($\mu$-barrier), and mixing between the core and the envelope is
inhibited. Exactly when this inhibition becomes important depends
on the initial angular momentum of the star. The amount of mixing
that occurs between the core and the envelope is affected accordingly.
Mixing inside the envelope also increases for larger initial angular
momentum, since the dominant mixing process, Eddington-Sweet
circulation, has an efficiency that increases as the square of the
stellar rotation rate.
As the evolution of the star proceeds to later stages, the time-scale
for rotationally induced mixing becomes too long in comparison to the
evolutionary time-scales to constitute an important source of
large-scale mixing. Also, the mixing is not able to dissolve the
molecular weight barrier which forms in the core during central helium
burning. In general, rotationally induced mixing does not strongly
affect the stellar structure after central helium ignition. The
evolution of the star from this point until core collapse is similar
to that of a non-rotating star of same structure at this time, except
for the differences in the nucleosynthesis discussed below.
For models where rotationally induced mixing is assumed to be
insensitive to gradients in the mean molecular weight, no
$\mu$-barrier inhibits the mixing. This affects the mixing between
the core and the envelope during central hydrogen burning. The
$\mu$-barrier in the superadiabatic part of the core during central
helium burning is eroded by shear instabilities. Consequently, the
convective core can grow unhindered. As a result, the helium cores
are more massive, corresponding to non-rotating stars with about
$25\,\%$ higher initial mass. Inside this helium core, the {CO} core
is also larger than that of a non-rotating star with same helium core
mass. Towards the end of central helium burning, fresh helium is
mixed into the convective core both by the continuing growth of this
core and by rotationally induced mixing. The fresh helium
preferentially converts carbon into oxygen instead of producing new
carbon by the triple-alpha process. This reduces the carbon abundance
in the core. Except for this, the effect of rotationally induced
mixing is small after helium ignition for the reason outlined above.
In particular, also in this case, the hydrogen burning shell
constitutes an efficient barrier for mixing processes --- indeed even
more efficient, because the core rotates slower as in the case where
$\mu$-gradients were considered (see below). A consequence of the
enlarged cores is that the limit on the initial stellar mass for core
collapse supernovae is somewhat smaller for higher initial rotation
rates.
\subsection{Angular momentum}
At central hydrogen ignition, the stars establish almost uniform
rotation. If a molecular weight barrier forms as hydrogen burning
progresses, angular momentum is trapped inside the core and
differential rotation results, with up to a factor of $\sim3$
variation in the rotation rate between the core and the envelope. If
$\mu$-barriers are unimportant for the rotationally induced mixing,
the stars stay close to rigid rotation until the end of central
hydrogen burning. Since this barrier forms later in the faster
rotators, stars having different initial rotation rates may end up
with similar specific angular momenta in the core at the end of
central hydrogen burning
(\FigCCc{D}{m-wjJ53-E15AB}{C}{m-wjJ53-GF15B}{D}). Due to angular
momentum transport during core helium burning, they may become even
more similar in the pre-collapse stage ({\Fig{jj15B}}). Some angular
momentum gets lost from the core during the restructuring that occurs
after core hydrogen exhaustion, but during central helium burning the
hydrogen-burning shell constitutes an efficient barrier that inhibits
the transport of angular momentum out of the core. Even so, the
average angular momentum of the core may decrease somewhat, since it
grows into regions with lower specific angular momentum on top of it.
The helium core itself stays close to uniform rotation
(\FigBb{A}{m-wjJ53-E15AB}{B}). During central helium burning
rotationally induced mixing processes already become slow compared to
the evolution, and after core helium exhaustion they do not cause any
relevant transport of angular momentum.
Only convective processes are rapid enough to notably redistribute
angular momentum during the late stages of stellar evolution. Within
the assumptions made, rigid rotation results in convective regions,
transporting angular momentum from their bottom to their top.
Subsequent phases of convective central and shell burning stages give
some outward transport of angular momentum inside the carbon-oxygen
core. Since none of the convective shells penetrates through the
outer boundary of the {CO} core, the angular momentum remains trapped
inside (\FigCCc{F}{m-wjJ53-E15AB}{E}{m-wjJ53-GF15B}{F}). The outer
boundary of the helium core constitutes a similar barrier.
This has interesting consequence for the final angular momentum in the
core. First, the different convective burning shells leave their
fingerprint not only in the chemical composition, but also in the
angular momentum distribution: a spiky profile results at the onset of
core collapse (\Fig{jj15B}). The high peaks correspond to the upper
edges of the most recently active convection zones and the deep
valleys to their bottoms. Shells of similar composition tend to
rotate almost rigidly. Second, even the slowest rotating core of the
Type~II supernova progenitor stars considered here would result in a
neutron star rotating close to break-up if angular momentum were
conserved during the collapse. This is not necessarily in
contradiction with observations of young neutron stars in supernova
remnants, even though the fastest of these rotates much slower. These
pulsars are already hundreds of years old, and recent theoretical
investigations of hot, newly born neutron stars indicate they may spin
down to the observed rotation rates within about a year by emitting
gravitational waves {\Cite{LOM98}}. The electro-magnetic radiation
emanating from pulsars is trapped inside the supernova ejecta during
that time, but the gravitational radiation of these very young neutron
stars might become detectable in the future {\Cite{owe98}}.
\acknowledgments
We are grateful to T.A.~Weaver for helpful discussions and aid with
KEPLER. This work was supported by the Deutsche Forschungsgemeinschaft
through grants La~587/15 and La~587/16, by the the National
Science Foundation (AST 97-31569), and by the Alexander von Humboldt
Foundation. AH was, in part, supported by a ``Doktorandenstipendium
aus Mitteln des 2.~Hochschulprogramms''.
|
\section{Introduction}
\bigskip
The possibility of analyzing narrow $s$-channel
resonances is considered to be one of the most important strengths
of muon colliders, at present under consideration by different laboratories
\cite{mumucolliders}.
Most of the content of this note is indeed addressed to physics
of particular relevance to future muon colliders
(for general reviews see \cite{physics}). The importance of
analyzing accurately an $s$-channel resonance at a lepton collider and the
physical interest to be assigned to the extracted information are, of course,
not new subjects. Much literature has in fact been devoted to
the problem of the $Z$ resonance shape at electron-positron colliders and to
the implications of its accurate measurement. Many of the special
issues that arise when studying a narrow $s$-channel resonance,
such as a light SM Higgs boson, have also been considered \cite{bbgh}.
However, these latter studies assumed that the beam energy profile
is perfectly known. If a resonance is so narrow that
its width is smaller than or comparable to the beam energy spread,
uncertainty in the beam's energy profile can introduce substantial errors
in the experimental determinations of the resonance parameters.
One of the main goals of the present paper is to assess these errors.
To this end, we shall present a detailed discussion of
methods for studying a narrow $s$-channel resonance or systems of nearly
degenerate resonances at a lepton collider, and consider applications to
particular cases. We carefully analyze the errors
in the measured resonance parameters (such as the total
width and partial widths) that arise from the uncertainty
in the energy spread of the beam. Procedures for reducing this
source of errors are studied and experimental observables with
minimal sensitivity to beam energy spread are emphasized.
Our study will assume that the beam energy spread is independently
measured, as will normally be possible. Expectations and procedures
for this determination at a muon collider will be noted.
Although the parameters of a narrow $s$-channel resonance can be
determined with minimal sensitivity to beam energy
spread by measuring the total Breit-Wigner area, the peak total cross
section and the cross sections in different final state channels,
these measurements are not all easily performed with good
statistical accuracy. Generally, a more
effective method for determining resonance parameters is
to scan the resonance using specific on- and off-resonance energy
settings. In this paper, we also discuss the
possibly superior new method introduced in Ref.~\cite{ratio}
in which one operates the collider always with center of mass
energy equal to the resonance mass but uses two different beam energy
spreads: one smaller than the resonance width and
one larger. All of these
possibilities will be compared. The most immediate application
is to the study of a light SM-Higgs, where, as already known
\cite{bbgh,gunmumu,higgsmumu}, accurate
measurements at a muon collider might make it possible
to distinguish a standard model Higgs from the lightest Higgs of a
supersymmetric model. We shall also analyze in detail resonant production
of the lightest pseudogoldstone boson of dynamical electroweak symmetry
breaking models. Following that, we consider two almost coincident
vector resonances of the Degenerate BESS model \cite{DBESS}
when their mass splitting is larger than the beam energy spread.
We shall then examine the case of two nearby resonances with
mass splitting much smaller than their average mass, assuming that
the energy spread of the beam is of the same order as the mass splitting.
Different possible procedures for
reducing the errors in the determination of the physical parameters
will be examined. As an
application we shall discuss such a situation in Degenerate BESS.
In Section 2 we present the analysis for a narrow resonance and in Section
3 the applications we have mentioned for this case. Section 4 gives the
analysis for nearly degenerate resonances, and Section 5 a corresponding
application.
\bigskip
\section{Analysis for a narrow resonance}
\label{general}
\bigskip
In this Section we will discuss several ways of determining the
parameters of a narrow resonance at a lepton collider.
For a given decay channel, an $s$-channel resonance can be described by three
parameters: the mass, the total width and the peak cross section.
These quantities can only be accurately measured if the beam parameters
are very well known. Measurement of the mass requires
precise knowledge of the absolute energy value.
Measurement of the width and cross section requires excellent
knowledge of the energy spread of the beam and the ability to determine
the difference between two beam energy settings with great precision.
All of these beam parameters must be known with extraordinary accuracy
in order to study a very narrow resonance. For example, consider
a light SM Higgs boson. In \cite{bbgh} it was found that
the beam energy $E_{\rm beam}$ must be known to better than 1 part in $10^6$
and the beam energy spread $\Delta E_{\rm beam}$
must be smaller than 1 part in $10^4$
in order to scan the resonance and determine its width and other
parameters. Below, we shall find that errors
in these parameters due to uncertainty
in $\Delta E_{\rm beam}$ will only be small compared to
statistical errors associated with the measurements of a typical cross section
(equivalent to a rate in a particular final state channel) if the error in the
measurement of $\Delta E_{\rm beam}$ is smaller than $\sim 1\%$.
The independent measurement of $\Delta E_{\rm beam}$ can have both
statistical and systematic error. Let us first consider the
impact of statistical
errors. We will argue that statistical uncertainties in $\Delta E_{\rm beam}$
(and also $E_{\rm beam}$ itself) should not be significant.
Consider measurement of some resonance parameter $p$ through a series
of observations over a large number ($S$) of spills. In a year of
operation at a muon collider there will be something like $S=10^8$ spills.
Each spill ({\it i.e.}\ each muon bunch) will contribute
to the measurement of $p$, but the value of $p$ interpreted
from the cross section $\sigma$ observed for a given spill will be uncertain
by an amount $\delta p$. This uncertainty arises (a)
because of the limited number of events $N_i$ accumulated during the spill
and (b) due to uncertainties in
$E_{\rm beam}$ and $\Delta E_{\rm beam}$ for that spill.
We have, using the short hand notation of $E_b$ for $E_{\rm beam}$,
\begin{equation}
\left[\delta p\over p\right]_i=\left[
{ c^2\over N_i}+ \left(a
{\delta E_b\over E_b}\right)_i^2 +
\left(b
{\delta \Delta E_b\over \Delta E_b}\right)_i^2
\right]^{1/2}\,,
\label{erroreq}
\end{equation}
where $c={d\ln p\over d \ln \sigma}$,
$a={d\ln p\over d\ln E_b}$ and $b={d\ln p\over d\ln \Delta E_b}$.
$E_b$ and $\Delta E_b$ are measured
by looking at oscillations in the spectra of the muon decay products
arising due to spin precession of the muons in the bunch
during the course of a very large number of turns
around the ring. These measurements are completely independent
of the measured rate(s) in question. The fractional
statistical errors ${\delta E_b\over E_b}$
and ${\delta \Delta E_b\over \Delta E_b}$ for each spill are expected
to be small as we shall summarize below. As a result, the contribution
from the ${c^2\over N_i}$ term in Eq.~(\ref{erroreq}) will normally
be much larger than that from the $\delta E_b$ and $\delta\Delta E_b$
terms unless the coefficients $a$ and $b$ are very large. We will
later learn that $a$ and $b$ will always be small under circumstances
in which systematic errors in $E_b$ and $\Delta E_b$ do not
badly distort the parameter measurement. To the extent that this
is true, the
statistical error $\delta p/p$ after a large number of spills,
computed as
\begin{equation}
{\delta p \over p}=\left[
\sum_i {1\over \left({\delta p\over p}\right)_i^2}\right]^{-1/2}\,,
\label{finalerror}
\end{equation}
will be dominated by the usual $c/\sqrt{\sum_i N_i}$ term. Note that if
this term were altogether absent, then ${\delta p\over p}$ would
be proportional to $1/\sqrt S$ times
the per spill error from $\delta E_b$ and $\delta \Delta E_b$.
Thus, even if this latter were quite substantial, it would be strongly
suppressed after accumulating data over $10^8$ spills.
Since $E_{\rm beam}$ and $\Delta E_{\rm beam}$
are expected to vary somewhat from spill to spill, what is important
is the statistical error with which they can be measured in a given spill.
This has been discussed in
\cite{spread,raja,reportusa,reporteurope}.
Very roughly, the frequency of
oscillation in the signal of secondary positrons from
the muon decays (which signal is sensitive to the precession
of the naturally-present polarization of the muons)
determines $E_{\rm beam}$ and the decay with time of the oscillation signal
amplitude determines $\Delta E_{\rm beam}/E_{\rm beam}$.
At a resonance mass of $100~{\rm GeV}$,~\footnote{The precise
figures given here are from \cite{reporteurope}.} a beam energy spread
of $\Delta E_{\rm beam}/E_{\rm beam}\sim 3\times 10^{-5}$ can be achieved
by an appropriate machine design
while maintaining adequate yearly luminosity ($\geq 0.1~{\rm fb}^{-1}$).
Further, in this case, {\it for each muon spill} ({{\it i.e.}\ each
muon bunch}), one can determine the beam
energy itself to 1 part in $10^7$ (5 keV) and measure
the actual magnitude of $\Delta E_{\rm beam}/E_{\rm beam}$ with
an accuracy of roughly $1.67\%$.
For $\Delta E_{\rm beam}/E_{\rm beam}\sim 10^{-3}$, as might
be useful for a much broader resonance than the SM Higgs boson
and as would allow for substantially larger yearly integrated luminosity
($\geq 1~{\rm fb}^{-1}$), $E_{\rm beam}$ can be measured to 2 parts in $10^{6}$ (100 keV)
and $\Delta E_{\rm beam}/E_{\rm beam}$ can be measured with
an accuracy of roughly $0.2\%$. As described above, this level
of statistical error for $E_{\rm beam}$ and $\Delta E_{\rm beam}/E_{\rm beam}$
will result in only a tiny $\delta p/p$ error for a given resonance
parameter unless the coefficients $a$ and $b$ of Eq.~(\ref{erroreq})
are very large. Thus, in what follows, we will analyze the parameter
errors introduced by an uncertainty in $\Delta E_{\rm beam}/E_{\rm beam}$
that is assumed to be systematic in nature.
The level of such uncertainty is not well understood at the moment.
The energy spread will be affected by possible time dependence of other
beam parameters, such as emmittance, by time-dependent backgrounds to
the precession measurements, and by other sources of depolarization.
Absent a detailed design for the machine and the polarimeter,
\cite{reporteurope} states that it is ``quite certain'' that the energy
spread can be known with relative systematic error better than $\sim 1\%$,
even if $\Delta E_{\rm beam}/E_{\rm beam}=3\cdot 10^{-5}$.
Given the precision with which the beam energy can
be determined, it is clear that the mass of a resonance
will be precisely known.
We will not consider here the statistical
errors in the measurements of the relevant cross sections,
but these will later become important when choosing among
the different procedures that we will discuss.
Our main focus will be on
the errors in the determination of branching ratios and
total widths induced by a systematic uncertainty in the energy spread of
the beam. Therefore, we will discuss the possibility of
reducing such errors, and also discuss measurements which are
independent of the energy spread. As already noted,
these errors are expected to be important only
when the width of the resonance is smaller than or comparable
to the energy spread of the beam.
We assume that the resonance is well described by a
Breit-Wigner shape. For a resonance $R$ of spin $j$
produced in the $s$-channel and decaying into a
given final state $F$, one has
\begin{equation}\label{BW}
\sigma^{F}(E)=4\pi(2 j+1)\frac{\Gamma(R\to\ell^+\ell^-)
\Gamma(R\to F)}{(E^2-M^2)^2+M^2\Gamma^2}
\end{equation}
where $M$ and $\Gamma$ are the mass and the width of the
resonance, and $E=\sqrt{s}$. We will work in the narrow
width approximation and therefore neglect the running of the
width. We will consider also the total production cross section
\begin{equation}\label{BWT}
\sigma(E)=4\pi(2 j+1)\frac{\Gamma(R\to\ell^+\ell^-)
\Gamma}{(E^2-M^2)^2+M^2\Gamma^2}
\end{equation}
We assume that in the absence of bremsstrahlung
the lepton beams have a Gaussian beam energy spread specified
by $\Delta E_{\rm beam}/E_{\rm beam}=0.01\,R(\%)$, leading to
a spread in the center of mass energy given by
\begin{equation}\label{spread}
\sigma_E={0.01\,R(\%)\over\sqrt 2}\,E\sim 0.007\,R(\%)\,E\,.
\end{equation}
In the absence of bremsstrahlung, the
energy probability distribution $f(E)$ can be written in the scaling form
\begin{equation}\label{scaling}
f(E)=\frac 1{\sigma_E} g\left(\frac {E-E_0}{\sigma_E}\right)\,,
\end{equation}
where $E_0$ is the central energy setting and
\begin{equation}\label{normalization}
\int f(E)\,dE=1\,.
\end{equation}
Accurate measurements of a narrow resonance require an accurate knowledge
of both $\sigma_E$ and of the shape function $g$.
We will explore the systematic uncertainties associated with errors
in determining $\sigma_E$, assuming that the shape function $g$
is known to be of a Gaussian form.
After including bremsstrahlung (at an $e^+e^-$ collider, beamstrahlung
must also be included) it is no longer possible to write
the full energy distribution in a scaling form; bremsstrahlung has
intrinsic knowledge of the mass of the lepton, which in leading
order enters in the form $\log(E/m_\ell)$.
Since the bremsstrahlung distribution is completely known it may be
convoluted with the scaling form of Eq.~(\ref{scaling})
to obtain a final energy distribution which depends upon both $\sigma_E$
and the (very accurately known) machine energy setting.
Some useful figures illustrating the effects of bremsstrahlung at
a muon collider are Figs. 30, 31 and 32 in Appendix A of \cite{bbgh}.
One typically finds that the peak luminosity at the central beam energy, $E_0$,
is reduced to about 60-80\% of the value in the absence of
bremsstrahlung, and that $d{\cal L}/dE$ falls below
$10^{-3}\left.d{\cal L}/dE\right|_{E=E_0}$ at $E\sim 0.9 E_0$.
We will discuss results at a muon collider
both before and after including bremsstrahlung, and show that
the errors introduced by uncertainty in the beam energy spread
can be very adequately assessed without including bremsstrahlung.
The measured cross sections are obtained from the convolution of
the theoretical cross sections [see (\ref{BW})] with the energy
distribution:
\begin{equation}\label{convolution}
\sigma_c^{F}(E)=4\pi(2j+1) \Gamma(R\to\ell^+\ell^-)
\Gamma(R\to F) h(\Gamma,\sigma_E,E)
\end{equation}
where
\begin{equation}\label{convolution2}
h(\Gamma,\sigma_E,E)=\int \frac{f(E-
E')}{({E'}^2-M^2)^2+M^2\Gamma^2}d E'
\end{equation}
From these equations we can immediately draw two useful
consequences:
\begin{itemize}
\item the ratio of $\sigma_c^{F}$ to
the production cross section is given by
\begin{equation}\label{ratio}
\frac{\sigma_c^{F}}{\sigma_c}=B_{F},~~~B_{F}=B(R\to F)
\end{equation}
and does not depend on $\sigma_E$;
\item due to the
normalization condition (\ref{normalization}), the integral of
$\sigma_c^{F}$ is independent of $\sigma_E$
\begin{equation}\label{integral}
\int\sigma_c^{F}(E)dE=\int\sigma^{F}(E)dE \,.
\end{equation}
Of course, one must be very certain that the integral over $E$ covers
all of the resonance and all of the beam spread, including the
low-energy bremsstrahlung tail.
\end{itemize}
Let us now consider the narrow resonance approximation. In
this case, we can write
\begin{equation}\label{NRA}
M^2\sigma^{F}(E)=\pi(2 j+1)B_{\ell^+\ell^-} B_{F}
\frac{\gamma^2}{x^2+\gamma^2/4}
\end{equation}
where we have scaled the energy and the total width in terms of
the energy spread evaluated at the peak (here we take
$\sigma_E\approx \sigma_M$)
\begin{equation}\label{newvariables}
x=\frac{E-M}{\sigma_M},~~~~\gamma=\frac{\Gamma}{\sigma_M}
\end{equation}
The convolution can also be written in terms of these variables,
with the result (see Eq. (\ref{scaling}))
\begin{eqnarray}\label{scaled-conv}
M^2\sigma_c^{F}(E)&=&\pi(2j+1)B_{\ell^+\ell^-}B_{F}
\int g(x-x')\frac{\gamma^2}{{x'}^2+\gamma^2/4}dx'\nonumber\\&\equiv
&B_{\ell^+\ell^-}B_{F}\,\Phi(x,\gamma)
\end{eqnarray}
In particular, for the production cross section, we get
\begin{equation}\label{scaled-conv3}
M^2\sigma_c(E)=B_{\ell^+\ell^-}\,\Phi(x,\gamma)
\end{equation}
Some simple results that are valid in the narrow resonance approximation
are the following:
\begin{itemize}
\item If $\Gamma\gg\sigma_M$ (and $\Gamma$ is also large
compared to bremsstrahlung energy spread), then
\begin{equation}
\sigma_c(E)={4\pi B_{\ell^+\ell^-}\over M^2}{\Gamma^2M^2\over
(E^2-M^2)^2+\Gamma^2M^2}\,.
\label{largegam}
\end{equation}
A measurement of the peak cross section (summed over all final state modes)
gives a direct measurement of $B_{\ell^+\ell^-}$. In addition,
the resonance shape can be scanned by using a series
of different energy settings for the machine and $\Gamma$ can
then be determined.
\item If $\Gamma\ll\sigma_M$, then
\begin{equation}
\sigma_c(E)={4\pi B_{\ell^+\ell^-} \over M^2}{\Gamma \sqrt\pi\over 2\sqrt
2\sigma_M}\exp\left[{-(E-M)^2\over 2\sigma_M^2}\right]\,,
\label{smallgam}
\end{equation}
where we have assumed a Gaussian form for the beam energy spread. (We also
neglected bremsstrahlung, but this could be easily included.)
In this case, the peak cross section determines $B_{\ell^+\ell^-}\cdot\Gamma$
{\it provided} $\sigma_M$ is accurately known.
$\sigma_M$ could in turn be measured by changing $E$ relative to $M$.
However, statistical accuracy may be poor because
cross sections are suppressed by $\Gamma/\sigma_M$ in comparison
to those obtained if $\Gamma\geq\sigma_M$.
\end{itemize}
When scanning a resonance,
the above limits make it clear that the cross section is largest
and that systematic errors associated with $\sigma_M$ are smallest
when $\sigma_M$ is as small compared to $\Gamma$ as possible.
However, for a sufficiently narrow resonance, the luminosity
reduction associated with achieving $\sigma_M$
values significantly smaller than $\Gamma$ is often so great
that statistical uncertainties become large. Thus, we are often
faced with a situation in which $\Gamma$ is comparable
to or only a few times larger than $\sigma_M$.
(For example, this will be the case when studying a light SM Higgs boson
at a muon collider.) It is in this
situation that systematic errors in the resonance
parameters deriving from uncertainty in
$\sigma_M$ can be enhanced.
We now discuss four ways of extracting $B_{\ell^+\ell^-}$,
$B_{\ell^+\ell^-}B_{F}$ in a specific final state channel
and $\Gamma$ from the data and assess the extent to which their
determination will be influenced by uncertainty associated
with an independent measurement of $\sigma_M$.
\begin{table}[t]
\caption{ Errors on $\Gamma$ and $B$
induced by $\Delta\sigma_M/\sigma_M= +0.05$ (upper lines) and
$\Delta\sigma_M/\sigma_M= -0.05$ (lower lines) evaluated by
choosing one observation at the peak and the other at
$E-M=k\,\Gamma$, for $k=1,2,3$. Bremsstrahlung is neglected.}
\begin{center}
\begin{tabular}{|c|l|l|l|l|l|l|l|l|c|}
\hline
&
\multicolumn{2}{|c|}{$1\,\Gamma$}&\multicolumn{2}{|c|}{$2\,\Gamma$}
&\multicolumn{2}{|c|}{$3\,\Gamma$}
\\
$\Gamma/\sigma_M$& $\Delta B/B$& $\Delta\Gamma/\Gamma$&$\Delta B/B$&
$\Delta\Gamma/\Gamma$&$\Delta B/B$&$\Delta\Gamma/\Gamma$
\\\hline
1 & ${+0.220}$ & ${-0.205}$ & ${+0.171}$
& ${-0.160}$ & ${+0.114}$ & ${-0.101}$\\
& ${-0.140}$ & ${+0.198}$ & ${-0.119}$
& ${+0.153}$ & ${-0.089}$ & ${+0.094}$\\
\hline
2 & ${+0.065}$ & ${-0.075}$ & ${+0.043}$
& ${-0.037}$ & ${+0.035}$ & ${-0.023}$\\
& ${-0.053}$ & ${+0.072}$ & ${-0.039}$
& ${+0.035}$ & ${-0.033}$ & ${+0.023}$\\
\hline
3 & ${+0.031}$ & ${-0.036}$ & ${+0.023}$
& ${-0.017}$ & ${+0.022}$ & ${-0.013}$\\
& ${-0.027}$ & ${+0.034}$ & ${-0.022}$
& ${+0.017}$ & ${-0.021}$ & ${+0.013}$\\
\hline
4 & ${+0.018}$ & ${-0.020}$ & ${+0.016}$
& ${-0.011}$ & ${+0.015}$ & ${-0.009}$\\
& ${-0.017}$ & ${+0.019}$ & ${-0.015}$
& ${+0.010}$ & ${-0.014}$ & ${+0.009}$\\
\hline
\end{tabular}
\end{center}
\label{nobremi}
\end{table}
\bigskip
\bigskip
\noindent {\bf{Scan of the resonance}} - If $\Gamma\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}}\sigma_M$,
one of the most direct ways to determine the parameters of
an $s$-channel resonance is through a three-point
scan \cite{bbgh}. In this method one measures the production
cross section $\sigma_c$ at three different energies. However,
the position of the peak is independent of the energy spread, and
we do not expect that uncertainty in the latter will induce an
error in the mass of the resonance (we have checked explicitly
that this is indeed the case). Therefore we will assume in the
following that the mass of the resonance is known with high
accuracy, and, as a consequence, we will take into account only
two scan points; one will be chosen at the peak and the
other one off the peak. We will shortly discuss how the
results depend on the position of this last point. The parameters
$(B,\Gamma)$ are then extracted by a two parameter
fit. Here, $B=B_{\ell^+\ell^-}$ if one is able to sum
over all final state modes, and $B=B_{\ell^+\ell^-}B_{F}$
if one focuses on a particular final state.
In principle this problem has a unique solution, by
deconvoluting the observed cross section. However, the error in
the determination of $\sigma_M$ (the energy spread at the peak of
the resonance) induces an error on the determination of the
parameters of the resonance. Assuming that the measured value of
$\sigma_c$ at a given energy $E$ is given by
$\sigma_c(E,B,\Gamma,\sigma_M)$ (here again we use the narrow
resonance approximation to put $\sigma_E\approx\sigma_M$), the
changes in the values of $B$ and $\Gamma$ that result if $\sigma_M$
is shifted by an amount $\Delta\sigma_M$
can be evaluated through the equation
\begin{equation}\label{condition1}
\sigma_c(E,B,\Gamma,\sigma_M)=
\sigma_c(E,B+\Delta B,\Gamma+\Delta\Gamma,\sigma_M+\Delta\sigma_M)
\end{equation}
or, from Eq. (\ref{scaled-conv3}),
\begin{equation}\label{condition2}
\Phi(x,\gamma)=
\Phi(x+\Delta x,\gamma+\Delta\gamma)
\left(1+\frac{\Delta B}{B}\right)
\end{equation}
where
\begin{eqnarray}\label{def}
\Delta x&=&x(\sigma_M+\Delta\sigma_M)-x(\sigma_M)\approx
-x\frac{\Delta\sigma_M}
{\sigma_M}\nonumber\\\Delta\gamma&=&\gamma(\sigma_M+\Delta\sigma_M)-\gamma(\sigma_M)
\approx \frac{\Gamma}{\sigma_M}\left(
\frac{\Delta\Gamma}{\Gamma}-\frac{\Delta\sigma_M}{\sigma_M}\right)
\end{eqnarray}
If we measure the cross section at two different energies, and assume that
$\Delta\sigma_M$ is the same at these two nearby energies
(as would be typical of a systematic error), we can easily
employ Eq. (\ref{condition2}) to determine the induced
fractional shifts $\Delta\Gamma/\Gamma$ and $\Delta B/B$. We fix one of the two
points at the peak ($x=0$), and we let the other one vary
between $E-M=\Gamma$ and $E-M=3\Gamma$. The induced
systematic errors are illustrated in
a series of Tables. In Tables \ref{nobremi} and \ref{nobremii},
we assume a Gaussian energy distribution without including bremsstrahlung
and tabulate errors induced by
$\Delta\sigma_M/\sigma_M=0.05$ and 0.01, respectively. As clear
from Eq. (\ref{condition2}), the fractional errors on $\Gamma$ and
$B$ depend only on the $E-M$ choice and on the reduced width $\gamma$,
but not on $B$.
\begin{table}[h]
\caption{Same as Table \protect\ref{nobremi} for
$\Delta\sigma_M/\sigma_M= +0.01$ (upper lines) and
$\Delta\sigma_M/\sigma_M= -0.01$ (lower lines).}
\begin{center}
\begin{tabular}{|c|l|l|l|l|l|l|l|l|c|}
\hline
&
\multicolumn{2}{|c|}{$1\,\Gamma$}&\multicolumn{2}{|c|}{$2\,\Gamma$}
&\multicolumn{2}{|c|}{$3\,\Gamma$}
\\
$\Gamma/\sigma_M$& $\Delta B/B$& $\Delta\Gamma/\Gamma$&$\Delta B/B$&
$\Delta\Gamma/\Gamma$&$\Delta B/B$&$\Delta\Gamma/\Gamma$
\\\hline
1 & ${+0.036}$ & ${-0.040}$ & ${+0.029}$
& ${-0.031}$ & ${+0.020}$ & ${-0.020}$\\
& ${-0.033}$ & ${+0.040}$ & ${-0.027}$
& ${+0.031}$ & ${-0.019}$ & ${+0.019}$\\
\hline
2 & ${+0.012}$ & ${-0.015}$ & ${+0.008}$
& ${-0.007}$ & ${+0.007}$ & ${-0.004}$\\
& ${-0.011}$ & ${+0.015}$ & ${-0.008}$
& ${+0.007}$ & ${-0.007}$ & ${+0.004}$\\
\hline
3 & ${+0.006}$ & ${-0.007}$ & ${+0.005}$
& ${-0.003}$ & ${+0.004}$ & ${-0.003}$\\
& ${-0.006}$ & ${+0.007}$ & ${-0.004}$
& ${+0.003}$ & ${-0.004}$ & ${+0.003}$\\
\hline
4 & ${+0.004}$ & ${-0.004}$ & ${+0.003}$
& ${-0.002}$ & ${+0.003}$ & ${-0.002}$\\
& ${-0.003}$ & ${+0.004}$ & ${-0.003}$
& ${+0.002}$ & ${-0.003}$ & ${+0.002}$\\
\hline
\end{tabular}
\end{center}
\label{nobremii}
\end{table}
As expected from our earlier discussion, the fractional errors become
smaller for larger $\Gamma/\sigma_M$.
The fractional errors are also decreased by moving the
second point of the scan away from the peak (see Tables \ref{nobremi}
and \ref{nobremii}).
On the other hand the statistical errors for $\sigma_c$ are increasing as
\begin{equation}\label{ratio2}
\sigma_c(E)/\sigma_c(M) =\frac{\Phi(x,\gamma)}{\Phi(0,\gamma)}
\end{equation}
becomes smaller. (Due to presence of background, the decrease
is not simply proportional to $\sqrt{\sigma_c(M)/\sigma_c(E)}$.)
By fixing the amount of luminosity that one can safely lose
without incurring substantial statistical error in $\sigma_c(E)$, and for
a given ratio $\Gamma/\sigma_M$, one can, by looking at Fig. \ref{fig1},
fix the second scan point. For instance, for $\Gamma=2\sigma_M$,
allowing for a loss in luminosity of 60\%, we see that the point to
be chosen is $x\approx 2$ corresponding to $E\approx M+\Gamma$.
By choosing the second point of the scan at $E=M+2\Gamma$, the
resulting behaviour of the systematic errors vs. $\gamma$ is illustrated in
Fig. \ref{fig2}. For a given uncertainty on the energy spread the errors
decrease for increasing $\gamma$. Therefore, as noted earlier,
one should employ
the smallest value of $\sigma_M$ consistent with having luminosity
large enough to give small statistical $\sigma_c$ errors. For
instance, at $\gamma=1$ the errors on the resonance parameters
are of order $2.5\div 3.5$ times $\Delta\sigma_M/\sigma_M$,
whereas for $\gamma=3$ they are down to $0.3\div 0.5$ times
$\Delta\sigma_M/\sigma_M$.
Of course, a full study of the
optimization of this procedure in a concrete case requires also
taking into account the third scan point. In fact, one has
the further freedom of moving the position of this point,
with corresponding changes in the errors. One can show that by taking
the third scan point in a symmetric position with respect to the
other off-peak point, one gets the same errors shown in
Tables \ref{nobremi} and \ref{nobremii} (in the absence of bremsstrahlung).
By taking an asymmetric configuration the results are
in between the ones obtained with the corresponding
symmetric configurations. For instance, by taking one point at
the peak, one at $E=M+2\Gamma$ and the third one at $E=M-\Gamma$,
we find, for $\gamma=1$ and $\Delta\sigma_M/\sigma_M=0.05$,
$\Delta\Gamma/\Gamma=-0.174$ and $\Delta B/B=+0.186$.
\begin{table}[t]
\caption{Errors for $\Gamma$ and $B$
induced by $\Delta\sigma_M/\sigma_M= +0.05$ for no bremsstrahlung
(first lines, from Table \ref{nobremi}), and after including bremsstrahlung
for $R=0.003\%$ (second lines),
$R=0.03\%$ (third lines) and $R=0.1\%$ (fourth lines).
The energy profile before bremsstrahlung is taken to be
a Gaussian specified by the $\sigma_M$
corresponding to a given $R$ taking $M=100~{\rm GeV}$.}
\begin{center}
\begin{tabular}{|c|l|l|l|l|l|l|l|l|c|}
\hline
&
\multicolumn{2}{|c|}{$1\,\Gamma$}&\multicolumn{2}{|c|}{$2\,\Gamma$}
&\multicolumn{2}{|c|}{$3\,\Gamma$}
\\
$\Gamma/\sigma_M$& $\Delta B/B$& $\Delta\Gamma/\Gamma$&$\Delta B/B$&
$\Delta\Gamma/\Gamma$&$\Delta B/B$&$\Delta\Gamma/\Gamma$
\\\hline
1 & ${+0.22}$ & ${-0.21}$ & ${+0.17}$ & ${-0.16}$ & ${+0.11}$ & ${-0.10}$\\
& ${+0.22}$ & ${-0.21}$ & ${+0.18}$ & ${-0.16}$ & ${+0.12}$ & ${-0.11}$\\
& ${+0.22}$ & ${-0.21}$ & ${+0.18}$ & ${-0.16}$ & ${+0.12}$ & ${-0.11}$\\
& ${+0.22}$ & ${-0.21}$ & ${+0.18}$ & ${-0.16}$ & ${+0.12}$ & ${-0.11}$\\
\hline
2 & ${+0.065}$ & ${-0.075}$ & ${+0.043}$ & ${-0.037}$ & ${+0.035}$ & ${-0.023}$\\
& ${+0.067}$ & ${-0.078}$ & ${+0.046}$ & ${-0.044}$ & ${+0.040}$ & ${-0.032}$\\
& ${+0.065}$ & ${-0.075}$ & ${+0.044}$ & ${-0.041}$ & ${+0.037}$ & ${-0.029}$\\
& ${+0.067}$ & ${-0.077}$ & ${+0.045}$ & ${-0.042}$ & ${+0.039}$ & ${-0.031}$\\
\hline
3 & ${+0.031}$ & ${-0.036}$ & ${+0.023}$ & ${-0.017}$ & ${+0.022}$ & ${-0.013}$\\
& ${+0.031}$ & ${-0.035}$ & ${+0.024}$ & ${-0.020}$ & ${+0.023}$ & ${-0.017}$\\
& ${+0.031}$ & ${-0.037}$ & ${+0.024}$ & ${-0.020}$ & ${+0.023}$ & ${-0.017}$\\
& ${+0.031}$ & ${-0.038}$ & ${+0.024}$ & ${-0.021}$ & ${+0.022}$ & ${-0.016}$\\
\hline
4 & ${+0.018}$ & ${-0.020}$ & ${+0.016}$ & ${-0.011}$ & ${+0.015}$ & ${-0.009}$\\
& ${+0.019}$ & ${-0.022}$ & ${+0.017}$ & ${-0.013}$ & ${+0.015}$ & ${-0.011}$\\
& ${+0.018}$ & ${-0.021}$ & ${+0.016}$ & ${-0.012}$ & ${+0.015}$ & ${-0.011}$\\
& ${+0.018}$ & ${-0.020}$ & ${+0.015}$ & ${-0.012}$ & ${+0.015}$ & ${-0.011}$\\
\hline
\end{tabular}
\end{center}
\label{brema}
\end{table}
\begin{table}[h]
\caption{Errors for $\Gamma$ and $B$
expressed as $\Delta \Gamma/\Gamma=c_\Gamma \Delta\sigma_M/\sigma_M$
and $\Delta B/B=c_B \Delta\sigma_M/\sigma_M$ in the limit of
small $\Delta\sigma_M/\sigma_M$. Same entry ordering and bremsstrahlung
assumptions as in Table~\ref{brema}.}
\begin{center}
\begin{tabular}{|c|l|l|l|l|l|l|l|l|c|}
\hline
&
\multicolumn{2}{|c|}{$1\,\Gamma$}&\multicolumn{2}{|c|}{$2\,\Gamma$}
&\multicolumn{2}{|c|}{$3\,\Gamma$}
\\
$\Gamma/\sigma_M$ & $~~~c_B$ & $~~~c_\Gamma$ & $~~~c_B$ & $~~~c_\Gamma$ &
$~~~c_B$ & $~~~c_\Gamma$
\\\hline
1 & 3.4 & $-$4.0 & 2.8 & $-$3.1 & 2.0 & $-$2.0 \\
& 3.6 & $-$4.3 & 2.8 & $-$3.2 & 2.1 & $-$2.1 \\
& 3.2 & $-$3.7 & 2.8 & $-$3.2 & 2.1 & $-$2.1 \\
& 3.4 & $-$3.9 & 2.8 & $-$3.2 & 2.1 & $-$2.2 \\
\hline
2 & 1.2 & $-$1.5 & 0.81 & $-$0.72 & 0.69 & $-$0.45 \\
& 1.2 & $-$1.5 & 0.85 & $-$0.81 & 0.72 & $-$0.59 \\
& 1.2 & $-$1.5 & 0.86 & $-$0.82 & 0.74 & $-$0.61 \\
& 1.2 & $-$1.5 & 0.85 & $-$0.82 & 0.74 & $-$0.61 \\
\hline
3 & 0.58 & $-$0.70 & 0.46 & $-$0.34 & 0.43 & $-$0.26 \\
& 0.62 & $-$0.76 & 0.48 & $-$0.40 & 0.43 & $-$0.32 \\
& 0.60 & $-$0.76 & 0.48 & $-$0.42 & 0.46 & $-$0.36 \\
& 0.57 & $-$0.71 & 0.48 & $-$0.42 & 0.44 & $-$0.36 \\
\hline
4 & 0.35 & $-$0.40 & 0.31 & $-$0.21 & 0.30 & $-$0.17 \\
& 0.39 & $-$0.45 & 0.30 & $-$0.24 & 0.30 & $-$0.26 \\
& 0.39 & $-$0.43 & 0.32 & $-$0.23 & 0.29 & $-$0.23 \\
& 0.35 & $-$0.40 & 0.29 & $-$0.24 & 0.28 & $-$0.20 \\
\hline
\end{tabular}
\end{center}
\label{derapprox}
\end{table}
Inclusion of bremsstrahlung might complicate the picture
developed above, since the induced errors could, in principle,
depend upon the resonance mass and the value
of $R$, even when holding $\Gamma/\sigma_M$ and $(E-M)/\sigma_M$ fixed.
To determine if the effects of bremsstrahlung are important
for determining systematic errors coming from uncertainty in $\sigma_M$,
we begin by considering again $\Delta \sigma_M/\sigma_M=+0.05$.
We adopt $M=100~{\rm GeV}$ and consider $R=0.003\%$, $0.03\%$, and $0.1\%$
at a muon collider. The resulting
errors are given in Table~\ref{brema} in comparison to
the no-bremsstrahlung results of the upper lines of Table~\ref{nobremi}.
The induced errors computed including bremsstrahlung are quite independent
of $R$ and (within the numerical errors of our programs) are
essentially the same as those computed in the absence
of bremsstrahlung. This can be further illustrated in the limit
of very small $\Delta\sigma_M/\sigma_M$. In this limit, one can simply
compute the errors induced in $B$ and $\Gamma$ by using a linear
expansion of Eq.~(\ref{condition1}) and solving the resulting matrix equation.
The results for $\Delta B/B$ and $\Delta \Gamma/\Gamma$ expressed
as a coefficient times $\Delta\sigma_M/\sigma_M$ are given in Table
\ref{derapprox} for four cases. The first case is that of no bremsstrahlung.
(Note that for $\Delta\sigma_M/\sigma_M=\pm 0.01$ the errors predicted
are quite close to those of Table \ref{nobremii}.) The other
three cases are for $M=100~{\rm GeV}$ and $R=0.003\%$, $0.03\%$, and $0.1\%$,
including bremsstrahlung. Once again, we see that the error coefficients
computed including bremsstrahlung do not depend much on $R$
and are essentially the same (within the numerical errors of our programs)
as the error coefficients computed without
including bremsstrahlung. Thus, in what follows we will
discuss errors from $\Delta\sigma_M/\sigma_M$ without including
the effects of bremsstrahlung. Once a particular resonance is
discovered, the effects of bremsstrahlung upon the results
presented here can be easily incorporated to whatever precision is required.
As regards $e^+e^-$ colliders, at which bremsstrahlung is a more substantial
effect, a study (similar to that above
for the $\mu^+\mu^-$ collider) shows that the Gaussian approximation
for studying systematic errors from $\Delta\sigma_M$ is
again reasonable. Although beamstrahlung is also important
at an $e^+e^-$ collider, we did not study its effects. However,
current $e^+e^-$ collider designs are such that the energy
spreading from beamstrahlung is typically comparable to or smaller than
that from bremsstrahlung.
An interesting general question regarding the scan procedure is how
to optimize the choice of $\sigma_M$ relative to $\Gamma$
so as to minimize the net statistical plus systematic error.
Let us consider using $k=1$ for the scan, assuming a resonance
mass of $M=100~{\rm GeV}$ and
that $\Delta\sigma_M/\sigma_M$ is small ($\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.02$),
so that the linear error expansion using coefficients $c_{\Gamma}$
and $c_B$ is valid.
A very rough parameterization for the induced error in $\Gamma$
when $0.2\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} \Gamma/\sigma_M\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 3$ and $\Delta\sigma_M/\sigma_M\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.02$
is
\begin{equation}
{\Delta\Gamma\over\Gamma}\sim3.3\left({\sigma_M\over\Gamma}\right)^{1.288}
{\Delta\sigma_M\over\sigma_M}\,.
\label{dgamform}
\end{equation}
The optimal choice of $R$
for determining the parameters of a given resonance depends
upon many factors: how the machine luminosity varies with $R$;
the variation of $\Delta\sigma_M/\sigma_M$ with $R$; the
variation of the $\Delta\sigma_M$-induced
errors as a function of $R$; the magnitude of the resonance width
(in particular, as compared to $\sigma_M$);
and the size of backgrounds in the important final states to
which the resonance decays. At a muon collider, $R$ in the range
$0.1\%\div0.15\%$ is the natural result and
allows maximal luminosity. Increasing $R$ above this range
does not result in significant luminosity increase; in contrast,
the luminosity declines rapidly as $R$ is decreased below this range.
A convenient parameterization for the luminosity of an $E=100~{\rm GeV}$
muon collider, valid for $0.003\%\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} R\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.12\%$, is:~\footnote{The
$L_{\rm year}$ parameterization interpolates the three
results of Table 5 in \cite{reporteurope} taken from \cite{reportusa}.}
\begin{equation}
L_{\rm year}=1.2~{\rm fb}^{-1}\,\left({R\over 0.12\%}\right)^{0.67362}\,.
\label{lform}
\end{equation}
The specific coefficient in Eq.~(\ref{lform}) represents the most
pessimistic estimate for the instantaneous luminosities.
On occasion we shall also discuss results assuming a factor of 10 larger
coefficient, which we shall refer to as the optimistic luminosity estimate.
We do not know what to expect for the variation of the systematic
error in $\sigma_M$ as a function of $R$. In order to understand
how the optimization might work, we have adopted on a purely adhoc basis
a form that mimics the per-spill {\it statistical} error:~\footnote{This
form for the statistical
$\Delta\sigma_M/\sigma_M$ error interpolates the results for
the two $R=0.003\%$ and $R=0.1\%$ cases given in Sec. 4.3.2
of \cite{reporteurope} using a power law form.}
\begin{equation}
{\Delta\sigma_M\over\sigma_M}=2\cdot 10^{-3}\,\left({R\over
0.12\%}\right)^{-0.57477}\,.
\label{dsigmaform}
\end{equation}
This form corresponds to a decreasing fractional systematic error
in $\sigma_M$ as $R$ increases, as seems a likely possibility.
Finally, it is useful to recall the value of $\sigma_M$ as a function of $R$:
$$\sigma_M=86~{\rm MeV}\,{M\over 100~{\rm GeV}}\,\left({R\over 0.12\%}\right)\,.$$
For $R=0.003\%$, the smallest $R$ that is likely
to be achievable, these quantities have the values $L_{\rm year}=0.1~{\rm fb}^{-1}$,
${\Delta\sigma_M/ \sigma_M}=0.0167$, and $\sigma_M\sim 2~{\rm MeV}$ (for
$M=100~{\rm GeV}$). The rough implications of these dependencies are as follows.
\begin{itemize}
\item
For a broad resonance, defined as one
with $\Gamma\gg0.001M$, one should operate the muon collider
at its natural $R$ value of order $0.12\%$. The $\Delta\sigma_M$-induced
errors will be very tiny, both because $\Gamma/\sigma_M$ is very large
and because $\Delta\sigma_M/\sigma_M$ should be small for such $R$.
A precision scan of the resonance is readily possible in this case.
Further, the resonance cross section is maximal for large $\Gamma/\sigma_M$
and both background and signal rates vary slowly as a function of $R$.
\item
For a resonance with $\Gamma<0.001M$, one will wish to reduce $R$
until $\Gamma\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} \sigma_M$. The primary reasons to avoid
$\Gamma<\sigma_M$ are to avoid the large $\Delta\sigma_M$-induced errors
summarized above and to maintain adequate sensitivity
to the resonance width. Keeping $\Gamma/\sigma_M>1$, one can
combine the above formulae with the predicted variation
of statistical error as a function of $R$ at fixed $L$ to
determine the optimal $R$ choice. To illustrate, we will focus on the
statistical error for $\Gamma$ obtained using the three-point
scan with measurements at $E=M,M\pm 2\sigma_M$. If $\Gamma/\sigma_M$ is
large (typically $>7\div 10$) the signal cross sections at the
three scan points are roughly independent of $R$ (but only for $R$ values
smaller than or not too much larger than
that corresponding to $\sigma_M$). Since
the background cross section is also independent
of $R$, statistical errors would not vary much with $R$
if the total integrated luminosity is held fixed.
More generally, one will find that, at fixed $L$,
$\Delta\Gamma/\Gamma\propto R^p$ with $p>0$. (This is a very
crude representation valid only for a limited range of $R$.)
The exact rate of increase depends upon both $\Gamma/\sigma_M$
and the background level. Such cases will be discussed shortly.
However, even if $p=0$, $\Delta\Gamma/\Gamma$ also varies
as $1/\sqrt L$, where the available $L$ is given in Eq.~(\ref{lform}).
Defining $f\equiv R/0.003\%$, the net result is that
we can write $(\Delta\Gamma/\Gamma)_{\rm stat.}=c_{\rm stat.} f^{p-0.337}$.
Meanwhile, if we insert the expressions for $\sigma_M$
and $\Delta\sigma_M/\sigma_M$ in the expression for
the $\Delta\sigma_M$-induced $\Delta\Gamma/\Gamma$, we find
$(\Delta\Gamma/\Gamma)_{\rm syst.}=c_{\rm syst.} f^{0.713}$, where
(for the $k=1$ three-point scan)
$c_{\rm syst.}\sim 0.06\left({2~{\rm MeV}\over\Gamma}\right)^{1.288}$
if $M=100~{\rm GeV}$; we have normalized $\Gamma$ to the approximate
width of a $100~{\rm GeV}$ SM Higgs boson.
If $p-0.337<0$, as would apply for $p\sim 0$, then,
if the statistical and systematic errors are added in quadrature,
the opposite signs of their $f$ exponents means that
there will be a minimum in the combined error as a function of $f$.
If we use as a reference the value $f_0$
such that $c_{\rm stat.}=c_{\rm syst.}$,
then one finds that the optimal $f$ is such that $f/f_0=R/R_0\sim 0.7$
when $p\sim 0$ (requiring $\Gamma/\sigma_M(R_0)\gg 1$).
In other words, one should choose a value of $R$ somewhat
smaller than that value which would yield equal statistical and systematic
errors.
However, large $\Gamma/\sigma_M$ cannot be achieved for
the very narrow Higgs and pseudogoldstone bosons discussed
later in the paper. Even
for the larger $M_H$ and $M_{\pzero}$ values of possible interest for this
optimization discussion, we have, at best,
$\Gamma/\sigma_M(R=0.003\%)\sim 3\div 5$.
Starting from $R=0.003\%$, one finds that, at fixed total $L$,
$\Delta\Gamma/\Gamma$ increases quite rapidly with increasing $R$.
As two examples, $\Gamma/\sigma_M(R=0.003\%)\sim 3$ ($\sim 5$)
for a SM-like Higgs boson with $M_H\sim 140~{\rm GeV}$ ($\sim 150~{\rm GeV}$).
In these two cases,
$(\Delta\Gamma/\Gamma)_{\rm stat.}$ increases by a factor $\sim \sqrt 10$
($\sim \sqrt 2$)
as $R$ increases from $0.003\%$ to $0.01\%$. Representing this
increase using a power law (it is actually somewhat faster than
a power law), we can roughly write
$(\Delta\Gamma/\Gamma)_{\rm stat.}\sim c R^p$
at fixed $L$, with $p\sim 1$ ($p\sim 0.3$) for $M_H=140~{\rm GeV}$
($M_H\sim 150~{\rm GeV}$). (Note that for $R$ above the $0.003\%$ to $0.01\%$
range, the $p$ values are much bigger.)
From the analysis given in the preceding paragraph,
we immediately see that if $p>0.337$, the best
statistical and systematic errors will both be achieved by
taking $R$ as small as possible. In the SM Higgs case, $\Gamma/\sigma_M>5$,
corresponding to $M_H>150~{\rm GeV}$ would be required before
$p<0.337$ and there could
be some possible gain from increasing $R$. Similarly, in the case of the
narrow pseudogoldstone boson $P^0$
it is only at the very largest $M_{\pzero}=200~{\rm GeV}$ mass considered
that $p$ falls just slightly below $0.337$.
In practice, our a priori knowledge of $\Gamma$
(from the initial scan needed to pin
down the precise location of the $H$ or $P^0$) will be too imprecise
to allow for such optimization; one should plan on operating
the machine at $R=0.003\%$ if initial information indicates
that we are dealing with a Higgs or pseudogoldstone boson.
\end{itemize}
\bigskip
\noindent {\bf{Measuring the on-peak cross sections for
different \boldmath$\sigma_M$}} -
In this technique \cite{ratio} one presumes that $M$ is already quite well
known and that one has in hand a rough idea of the size of $\Gamma$.
(This is the likely result after the first rough scan used to locate
the resonance.) One then operates the collider at $E=M$
for two different values of $\sigma_M$ (spending perhaps a year or two
at each value). The results of Eqs.~(\ref{largegam}) and (\ref{smallgam})
show that if $\sigma_M^{\rm min}\ll\Gamma$ and $\sigma_M^{\rm max}\gg\Gamma$,
then $\sigma_c(\sigma_M^{\rm min})/\sigma_c(\sigma_M^{\rm max})={2\sqrt
2\sigma_M^{\rm max}\over \Gamma\sqrt\pi}$.
The systematic error in $\Gamma$ is
then given by
$(\Delta\Gamma/\Gamma)_{\rm syst.}
=\Delta\sigma_M^{\rm max}/ \sigma_M^{\rm max}$.
In practice, $\sigma_M^{\rm max}/\sigma_M^{\rm min}$
will be limited in size. If we define
$\sigma_M^{\rm central}=\sqrt{\sigma_M^{\rm max}\sigma_M^{\rm min}}$
(the geometric mean value) and
compute $r_c\equiv\sigma_c(\sigma_M^{\rm min})/\sigma_c(\sigma_M^{\rm max})=
\Phi(0,\gamma_{\rm min})/\Phi(0,\gamma_{\rm max})$
(where $\gamma_{\rm max,min}=\Gamma/\sigma_M^{\rm max,min}$) as
a function of $\Gamma$, our ability to measure $\Gamma$
in this way for any given value of
$\sigma_M^{\rm max}/\sigma_M^{\rm min}$ is determined
by the slope $|s|$ of $\ln[\Gamma/\sigma_M^{\rm central}]$
plotted as a function of $\ln[r_c]$.
The relevant plot from Ref.~\cite{ratio} appears in Fig.~\ref{dlgdlr}.
For a known $\sigma_M^{\rm central}$, the $|s|$
at any $\Gamma/\sigma_M^{\rm central}$ gives the relation
$(\Delta\Gamma/\Gamma)_{\rm stat.}=|s|(\Delta r_c/r_c)_{\rm stat.}$,
where $\Delta r_c/r_c$ is computed by combining the
fractional statistical errors for $\sigma_c(\sigma_M^{\rm min})$
and $\sigma_c(\sigma_M^{\rm max})$ in quadrature.
The point at which the magnitude of the slope, $|s|$,
is smallest indicates the point
at which a given fractional statistical error in the cross section ratio
will give the most accurate determination
(as measured by fractional error) of $\Gamma/\sigma_M^{\rm central}$.
We observe that $\Gamma/\sigma_M^{\rm central}\sim 2\div 3$ gives the
smallest $|s|$ (and hence smallest statistical error),
although $|s|$ at $\Gamma/\sigma_M^{\rm central}\sim 1$
is not that much larger. As expected,
the larger $\sigma_M^{\rm max}/\sigma_M^{\rm min}$, the smaller $|s|$
at any given $\Gamma/\sigma_M^{\rm central}$.
The systematic error in $\Gamma$ due to uncertainties in $\sigma_M^{\rm max}$
and $\sigma_M^{\rm min}$ is obtained from
\begin{equation}
{\Phi\left(0,{\Gamma\over \sigma_M^{\rm min}}\right)\over
\Phi\left(0,{\Gamma\over\sigma_M^{\rm max}}\right)}=
{\Phi\left(0,{\Gamma+\Delta\Gamma\over \sigma_M^{\rm
min}+\Delta\sigma_M^{\rm min}}\right)\over
\Phi\left(0,{\Gamma+\Delta\Gamma \over\sigma_M^{\rm max}
+\Delta\sigma_M^{\rm max}}\right)}\,.
\label{dgamratio}
\end{equation}
In most instances, the $\sigma_M^{\rm min}$ measurement is not
very sensitive to the precise value of $\sigma_M^{\rm min}$, in which case
the result is $(\Delta\Gamma/\Gamma)_{\rm syst.}
=\Delta\sigma_M^{\rm max}/\sigma_M^{\rm max}$.
More generally $(\Delta\Gamma/\Gamma)_{\rm syst.}=\Delta\sigma_M/\sigma_M$
to the extent that $\Delta\sigma_M/\sigma_M$ is the same at the different
$\sigma_M$ settings (as might apply for systematic errors of a certain type).
From this we see one clear advantage of this technique: the systematic
error in $\Gamma$ due to systematic error in $\sigma_M$ does not
increase with decreasing $\Gamma/\sigma_M^{\rm max,min}$. If statistical
errors for measuring the two $\sigma_c$'s can be kept small,
this is a clear advantage of the technique as compared to the scan
technique in which small $\Gamma/\sigma_M$ leads to large
$\Delta\sigma_M$-induced systematic error even if statistics
are excellent for all measurements. We will shortly discuss
issues related to statistical errors.
To continue our analysis of systematic errors,
let us note that, in the limit of very high statistics
for the cross section measurements ({\it i.e.}\ zero statistical error
for $r_c$), precise values of both
$\gamma_{\rm min}=\Gamma/\sigma_M^{\rm min}$ and
$\gamma_{\rm max}=\Gamma/\sigma_M^{\rm max}$ are obtained by the above procedure.
Further, $\sigma_c(\sigma_M^{\rm min})$ and $\sigma_c(\sigma_M^{\rm max})$
are given by $B\Phi(0,\gamma_{\rm min})$ and $B\Phi(0,\gamma_{\rm max})$,
respectively, where $B$ stands for
$B_{\ell^+\ell^-}$ or $B_{\ell^+\ell^-}B_F$, depending upon whether we are
looking at the total rate or the rate in some particular final channel $F$.
As a result, there is no systematic error in $B$
from uncertainty in $\sigma_M$, only statistical error
associated with the number of events observed. This is another advantage
of this technique.
Of course small systematic errors are not important
if statistical errors for the technique are not also small.
We summarize the considerations \cite{ratio}.
For the SM Higgs and the $P^0$ we found that it is best to use
$\sigma_M^{\rm min}$ corresponding to $R=0.003\%$ and
$\sigma_M^{\rm max}$ corresponding to $R=0.03\%$, so that
$\sigma_M^{\rm central}$ corresponds to $R=0.003\%\times\sqrt{10}$.
The parameterization for the variation of $L_{\rm year}$ given
in Eq.~(\ref{lform}) implies that $L_{\rm year}=0.1~{\rm fb}^{-1}$ ($0.47~{\rm fb}^{-1}$)
for $R=0.003\%$ ($0.03\%$). If, for example, $\Gamma/\sigma_M^{\rm central}=1$,
one finds $\sigma_c(\sigma_M^{\rm min})/\sigma_c(\sigma_M^{\rm max})=4.5$, implying that
the signal rate $S(\sigma_M)=L_{\rm year}(\sigma_M)\sigma_c(\sigma_M)$ is nearly
the same for $\sigma_M^{\rm max}$ as for $\sigma_M^{\rm min}$. However,
the background rate $B$ is proportional
to $L$ and thus $B/S$ is a factor of 4.7 times larger at $\sigma_M^{\rm max}$
than at $\sigma_M^{\rm min}$. Consequently, the statistical error in the measurement
of $\sigma_c(\sigma_M^{\rm max})$ will be worse than for
$\sigma_c(\sigma_M^{\rm min})$ for the same $S$.~\footnote{In the scan procedure,
there is a similar difficulty. There, $B/S$ is large
for the off-peak measurements.}
For a given running time at a given $\sigma_M$,
one must compute the channel-by-channel $S$ and $B$ rates,
compute the fractional error in $\sigma_c(\sigma_M)$
for each channel, and then combine all channels
to get the net $\sigma_c(\sigma_M)$ error.
This must be done for $\sigma_M=\sigma_M^{\rm min}$
and $\sigma_M=\sigma_M^{\rm max}$. One then computes the net
$r_c$ and net $\sigma_c$ errors as:
\begin{eqnarray}
{\Delta r_c\over r_c}&=&
\left\{\left[{\Delta \sigma_c(\sigma_M^{\rm min})\over\sigma_c(\sigma_M^{\rm min})}\right]^2
+\left[{\Delta \sigma_c(\sigma_M^{\rm max})\over\sigma_c(\sigma_M^{\rm max})}\right]^2
\right\}^{1/2}\,;
\label{rcerror}\cr
{\Delta\sigma_c\over\sigma_c}&=&
\left\{\left[{\Delta \sigma_c(\sigma_M^{\rm min})\over \sigma_c(\sigma_M^{\rm min})}\right]^{-2}
+\left[{\Delta \sigma_c(\sigma_M^{\rm max})\over\sigma_c(\sigma_M^{\rm max})}\right]^{-2}
\right\}^{-1/2}\,.
\label{sigcerror}
\end{eqnarray}
The ratio of running times at $\sigma_M^{\rm min}$ vs.
$\sigma_M^{\rm max}$ cannot be chosen so as to simultaneously
minimize the net $\Delta\sigma_c/\sigma_c$
and $\Delta r_c/r_c$. The former is minimized by running
only at $\sigma_M^{\rm min}$, while the latter is typically
minimized for $t(\sigma_M^{\rm min})/t(\sigma_M^{\rm max})\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$. For the SM Higgs,
a good compromise is to take $t(\sigma_M^{\rm min})/t(\sigma_M^{\rm max})=1$.
As demonstrated in the next section,
it turns out that for both the $P^0$ and the SM Higgs boson,
the ratio $L_{\rm year}(R=0.03\%)/L_{\rm year}(R=0.003\%)\sim 4.7$
and the predicted cross sections and backgrounds are such
that this technique is very competitive with the scan
technique as regards statistical errors for $\Gamma$.
Of course, for resonances (such as those of the Degenerate BESS model
considered later) that have fairly large widths, the normal scan
procedure can achieve superior results to the $r_c$-ratio technique.
This is because there will be little sacrifice in luminosity associated
with choosing an $R$ such that $\Gamma/\sigma_M$ is substantially
larger than 1. Measurements on the wings of the resonance will have
nearly the same statistical accuracy as measurements at the resonance peak.
\bigskip
\noindent{\bf{Measurement of $\Gamma$ using the ratio of the
$\ell^+\ell^-$ final state and total cross sections}} -
First of all let us discuss
the measurement of absolute and relative branching ratios. This is
possible by simply measuring the cross sections in different
final states, holding the energy fixed:
$\sigma_c^{F_1}/\sigma_c^{F_2}=B_{F_1}/B_{F_2}$
and $\sigma_c^{F}/\sigma_c=B_{F}$.
As apparent from Eq.~(\ref{ratio}), such cross section ratios do not depend on
$\sigma_M$. Therefore, we can measure
branching ratios and ratios of branching ratios
with no systematic error induced by the energy spread. Of course,
to measure $\sigma_c$, we must be able to sum over all final
states for which $B_{F}$ is substantial, taking into account
backgrounds and systematics associated with
correctly determining the relative normalizations
of different final states. This might not be easy, and could
even be impossible if the resonance has some effectively invisible decays
unless the branching ratio for invisible decays can be measured
in a different experimental setting (in particular, one
in which invisible decays can be effectively `tagged'
by producing the resonance in association with some other particle).
Given $B_{\ell^+\ell^-}$, the value
of $\sigma_c$ and Eq.~(\ref{scaled-conv3})
can then be used to determine $\Gamma$ if $\sigma_M$
is known. The error on $\Gamma$
induced by uncertainty in $\sigma_M$, given the absence of systematic error in
$B_{\ell^+\ell^-}$, is obtained from Eq. (\ref{condition2}) with $\Delta B=0$,
\begin{equation}\label{condition3}
\Phi(x,\gamma)=\Phi(x+\Delta x, \gamma+\Delta\gamma)\,.
\end{equation}
If we measure $\sigma_c$ and
$\sigma_c^{\ell^+\ell^-}$ at the resonance peak we have $\Delta x=0$ and
Eq.~(\ref{condition3}) then requires $\Delta\gamma=0$; in turn,
\begin{equation}
\Delta\gamma=0\longrightarrow \frac{\Delta\Gamma}{\Gamma}=\frac
{\Delta\sigma_M}{\sigma_M}\,.
\end{equation}
Therefore, the systematic error induced in $\Gamma$
does not depend on $\gamma$. On the other
hand, if we move away from the peak, the situation changes
drastically, as illustrated in Fig. \ref{fig3} which gives
$\Delta\Gamma/\Gamma$ vs. $\gamma$ for $\Delta\sigma_M/\sigma_M$
equal to 0.05. This figure can be trivially scaled for different
values of $\Delta\sigma_M/\sigma_M$. We see that above
$\gamma\approx 2$ the error decreases as one moves further away from the
resonance peak. For values below $\gamma\approx 2$ the situation is more
complicated, but for each choice of the energy there is a zero in
the error.~\footnote{A really precise computation
of the locations of these zeroes would need to include
the effects of bremsstrahlung.}
Therefore, by opportunely choosing the energy of the
measurement one could try to work in a region where the
systematic error is
very much reduced. For instance, if $\Gamma\approx 1.4\sigma_M$,
by taking $E=M+\Gamma$, one has $\Delta\Gamma=0$. Notice that the
locations of the zeroes do not depend on $\Delta\sigma_M$.
There are two potentially severe difficulties associated with
this technique. First, in many instances the $\ell^+\ell^-$
decay mode has a small branching ratio. If the resulting
event rate in the $\ell^+\ell^-$ final state is not large, the statistical
error for $\sigma_c^{\ell^+\ell^-}$ will be large. Statistics for $E$
significantly different from $M$ (as required to take advantage of
the zeroes discussed above) would be the first to become problematical.
Second, we have already noted that
measurement of $\sigma_c$ may be quite tricky. One must
correctly normalize different final state channels relative to one
another and assume that all final state channels
with substantial branching ratio are visible.
Regarding the latter, there is the alternative of measuring
the branching ratio for invisible final states
using other experimental situations/techniques and then making the appropriate
correction to compute the full $\sigma_c$ given
the contributions to $\sum_{F}\sigma_c^{F}$ that
can be measured in the $s$-channel setting.
Of course, in
some instances $B_{\ell^+\ell^-}$ will also be measured at other
machines or via other processes. Also in this case, one can immediately
determine $\Gamma$ given a measurement of $\sigma_c$. Alternatively,
if $B_{F}$ is also known for some final state $F$ (as well as
$B_{\ell^+\ell^-}$), $\sigma_c^F/(B_{\ell^+\ell^-}B_F)$ will
give a determination of $\Gamma$ with systematic uncertainty
from $\Delta \sigma_M$ as described above.
\bigskip
\noindent {\bf{Measurement of the Breit-Wigner area}} - As apparent from
Eq. (\ref{integral}), the energy integral of the cross section
does not depend on $\sigma_M$, and, in the narrow width
approximation, is given by
\begin{equation}\label{integral2}
\int\sigma_c(E)\,dE= 4\pi^2(2j+1)\,\Gamma\cdot\,B_{\ell^+\ell^-}
\end{equation}
Therefore, even if the energy spread is very poorly known,
so long as it does not change as one scans over the resonance one
could still measure the integral of the cross section, which is
proportional to the product $\Gamma\cdot B_{\ell^+\ell^-}$,
and obtain $B_{\ell^+\ell^-}$ from the ratio of the peak cross sections
$\sigma_c^{\ell^+\ell^-}/\sigma_c$
(or, possibly, in another experimental setting). Of course, as discussed
with regard to the previous procedure, the statistical error for
$\sigma_c^{\ell^+\ell^-}$ might be large
even when measured at the resonance peak. In addition,
the event rate substantially off-resonance (as compared
to both $\Gamma$ and $\sigma_M$) would be small
(while typical backgrounds would be essentially constant).
Thus, one can often only measure the cross section with
good statistical accuracy over a limited
portion of the full range.
If $\Gamma\gg\sigma_M$, and $\sigma_M$ is known ahead of time
(via spin-precession measurements or the like), this fact will
quickly become apparent after measuring the cross section at a few
energy settings with $E-M\gg\sigma_M$.
A reliable value for $\Gamma\cdot\,B_{\ell^+\ell^-}$
can normally be obtained so long as statistics are good at the peak.
The procedure is that based on Eq.~(\ref{largegam}).
One will measure $\sigma_c(E)$
for a set of points perhaps out to $E-M\sim \pm 2\Gamma$,
and the remainder of the integral will be determined using
the Breit-Wigner shape that would have been revealed by
the measurements made. Sensitivity to $\sigma_M$ will be
very minimal.
If $\sigma_M\gg\Gamma$, then the energy dependence of $\sigma_c(E)$
is determined by $\sigma_M$, see Eq.~(\ref{smallgam}),
or more generally by $g$ of Eq.~(\ref{scaling})
(bremsstrahlung should be included). Given a known value for $\sigma_M$
and a set of measured points, the remainder of the integral can be computed.
However, statistical errors are likely to be quite large because
the cross section is suppressed by the ratio $\Gamma/\sigma_M$.
Clearly, the most difficult case is that
in which the resonance is very narrow and
the smallest achievable $\sigma_M$ value is such
that $\Gamma\sim\sigma_M$.
Unless statistics remain good far off the peak,
which is not likely,
deconvolution of the effects of $\sigma_M$ and $\Gamma$ is required.
The simplest deconvolution procedures for known $\sigma_M$ are the
scan and ratio procedures outlined previously.
Thus, if $\Gamma\sim \sigma_M$, the three-point scan technique
and the technique of varying $\sigma_M$ while keeping $E=M$ are
the most efficient for determining the parameters of an
$s$-channel resonance so long as systematic uncertainty in the energy spread
of the beam is smaller than $1\%\div 2\%$.
(Of course, the appropriate strategy for exploring a resonance
would be quite different if $\sigma_M$ is not independently
measured with high statistical accuracy using the precession measurements.)
\bigskip
\section{Applications}
\bigskip
\noindent {\bf{SM-Higgs boson}} - In
this Section we will apply the first two methods of Section
\ref{general} to the study of a light SM-Higgs boson.
Consider first the three-point scan. By using
the results of Fig. \ref{fig2} one can easily determine the errors induced
by the uncertainty in the energy spread as a function of the Higgs
mass. For the evaluation of the Higgs width as a function of the
mass, we make use of the expressions given in Ref.
\cite{djouadi}. The decay channels we have considered are $b\bar
b$, $\tau^+\tau^-$, $c\bar c$, $gg$ and $WW^*$, $ZZ^*$ (one
of the two final bosons being virtual). The resulting total width
is given in Fig. \ref{fig4}. We have considered the interval $50\le
M_H({\rm GeV})\le 150$, but recall that, actually, there is an
experimental lower bound at $90\%$ C.L. of about 90 GeV
\cite{vancouver}. Below 110 GeV, the width of the Higgs
increases approximately linearly with the mass (aside from logarithmic
effects due to the running of the quark masses) which means that the
ratio $\Gamma_H/\sigma_M$ is approximately constant. By choosing
$R=0.003\%$ (see \cite{gunmumu}) we get $\Gamma_H/\sigma_M\approx
1$. From Fig. \ref{fig2}, we see that the $\Delta\sigma_M$-induced errors
in $B$ and in $\Gamma_H$ are
about 15\% and 2.5\% for $\Delta\sigma_M/\sigma_M=0.05$ and 0.01,
respectively, for a $k=2$ scan (as appropriate, given that
$\Gamma_H\sim\sigma_M$, for comparing
to the $E=M,~E=M\pm 2\sigma_M$ scan summarized below
that was used to estimate statistical errors).
Above 110 GeV, the systematic errors decrease rapidly due to
the fast increase of the width. Fig. \ref{fig5} summarizes
the $\Delta B/B$ and $\Delta \Gamma_H/\Gamma_H$ fractional
systematic errors as a function of $M_H$ for $R=0.003\%$ and $k=2$.
Notice that, at least in the region up to 110 GeV, it will be
mandatory to have values of $R$ of the order 0.003\%. For instance,
if we take $R=0.01\%$, corresponding to
$\Gamma_H/\sigma_M\approx 0.3$ (for $50\le M_H({\rm GeV})\le 110$), the
fractional $\Delta\sigma_M$-induced errors in $B$ and $\Gamma_H$
increase to about $14\%\div 16\%$ for $\Delta\sigma_M/\sigma_M=0.01$,
as can be seen from the dashed line of Fig. \ref{fig2}.
\begin{table}[h]
\caption[fake]{\baselineskip 0pt Percentage errors ($1\sigma$) for
$\sigma_cB(H\to b\overline b,WW^\star,ZZ^\star)$
(extracted from channel rates)
and $\Gamma_H$ for $s$-channel Higgs production at the muon collider
assuming beam energy resolution of $R=0.003\%$. Results are
presented for two integrated four-year luminosities:
$L=4~{\rm fb}^{-1}$ ($L=0.4~{\rm fb}^{-1}$). An optimized three-point scan
is employed using measurements at $E=M_H$,
$E=M_H+2\sigma_M$ and $E=M_H-2\sigma_M$, with luminosities of $L/5$,
$2L/5$ and $2L/5$, respectively. [For the cross section measurements,
this is equivalent to $L\sim 2~{\rm fb}^{-1}$ ($L=0.2~{\rm fb}^{-1}$) at the $E=M_H$ peak].
This table is taken from Ref.~\cite{gunmumu}. Efficiencies and cuts
are those employed in \cite{bbgh}. The effects of bremsstrahlung are
included.}
\small
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
Quantity & \multicolumn{4}{c|}{Errors} \\
\hline
\hline
{\bf Mass (GeV)} & {\bf 80} & {\bf $M_Z$} & {\bf 100} & {\bf 110} \\
\hline
$\sigma_cB(b\overline b)$ & $0.8\%(2.4\%)$ & $7\%(21\%)$ & $ 1.3\%(4\%)$ &
$ 1\%(3\%)$ \\
\hline
$\sigma_cB(WW^\star)$ & $-$ & $-$ & $ 10\%(32\%)$ &
$ 5\%(15\%)$ \\
\hline
$\sigma_cB(ZZ^\star)$ & $-$ & $-$ & $-$ &
$ 62\%(190\%)$ \\
\hline
$\Gamma_H$ & $ 3\%(10\%)$ & $ 25\%(78\%)$ & $ 10\%(30\%)$ &
$ 5\%(16\%)$ \\
\hline
\hline
{\bf Mass (GeV)} & {\bf 120} & {\bf 130} & {\bf 140} & {\bf 150} \\
\hline
$\sigma_cB(b\overline b)$ & $1\%(3\%)$ & $1.5\%(5\%)$ & $ 3\%(9\%)$ &
$ 9\%(28\%)$ \\
\hline
$\sigma_cB(WW^\star)$ & $3\%(10\%)$ & $2.5\%(8\%)$ & $ 2.3\%(7\%)$ &
$ 3\%(9\%)$ \\
\hline
$\sigma_cB(ZZ^\star)$ & $16\%(50\%)$ & $10\%(30\%)$ & $ 8\%(26\%)$ &
$ 11\%(34\%)$ \\
\hline
$\Gamma_H$ & $ 5\%(16\%)$ & $ 6\%(18\%)$ & $ 9\%(29\%)$ &
$ 34\%(105\%)$ \\
\hline
\end{tabular}
\end{center}
\label{fmcerrors}
\end{table}
It is interesting to compare these systematic errors to the statistical
errors. The analysis at a muon collider done in Ref. \cite{gunmumu}
gives statistical errors for a three-point scan using scan points
at $E=M,~E=M\pm 2\sigma_M$ and $R=0.003\%$, assuming
$L=4~{\rm fb}^{-1}$ or $L=0.4~{\rm fb}^{-1}$ total accumulated luminosity
(corresponding to 4 years of operation for optimistic or pessimistic,
respectively, instantaneous luminosity). The results
of that analysis are summarized in Table~\ref{fmcerrors}.
Except for $M_H\sim M_Z$, the $L=4~{\rm fb}^{-1}$
statistical error for measuring the total width would be of order
$3\%\div 10\%$ when $80\le M_H\le 140~{\rm GeV}$. Therefore, to avoid contaminating
this high precision measurement error with systematic
uncertainty from $\Delta\sigma_M$ we will certainly want
to have $\Delta\sigma_M/\sigma_M\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 1\%$.
If one adopts the $L=0.4~{\rm fb}^{-1}$
luminosity assumption (the benchmark value of
Refs.~\cite{reportusa,reporteurope}),
and if $m_H<130~{\rm GeV}$ and not near $M_Z$,
the statistical measurement error
for $\Gamma_H$ is in the 10\% to 20\% range. This means
that $\Delta\sigma_M/\sigma_M$ as large as $5\%$ would
be very undesirable.
Finally, we re-emphasize the fact that performing the
scan using larger $R$ than $R=0.003\%$ leads to larger
statistical errors until $M_H$ approaches the $WW$ decay
threshold and $\Gamma_H/\sigma_M(R=0.003\%)>5$. For lower $M_H$,
the $R=0.003\%$ results are the best
that can be achieved despite the smaller luminosity at $R=0.003\%$
as compared to higher $R$ values. For example, the error in $\Gamma_H$
for a given luminosity using $R=0.01\%$
can be read off from Fig.~13 of \cite{bbgh}. One finds that
$L(R=0.01\%)/L(R=0.003\%)=20,10,2$ is required in order that
the $\Gamma_H$ statistical errors for $R=0.01\%$ be equal to those
for $R=0.003\%$ at $M_H=130,140,150~{\rm GeV}$, respectively. Existing
machine designs are such that
$L_{\rm year}(R=0.01\%)/L_{\rm year}(R=0.003\%)=0.22~{\rm fb}^{-1}/0.1~{\rm fb}^{-1}=2.2$. Thus,
increasing $R$ would not improve the scan-procedure statistical errors
until $M_H>150~{\rm GeV}$. In addition, $\Delta\sigma_M$-induced
systematic errors always rise rapidly with increasing $R$. For
$M_H\leq 150~{\rm GeV}$, one should employ the smallest value of $R$ possible.
\begin{table}[h]
\caption[fake]{\baselineskip 0pt Percentage errors ($1\sigma$) for
$\sigma_cB(H\to b\overline b,WW^\star,ZZ^\star)$
(extracted from channel rates)
and $\Gamma_H$ for $s$-channel Higgs production at the muon collider.
Operation at $E=M_H$ is assumed and
the $r_c$-ratio technique is used for determining $\Gamma_H$.
Results are presented assuming accumulated luminosities of
$L=2~{\rm fb}^{-1}$ ($L=0.2~{\rm fb}^{-1}$) at $R=0.003\%$
and $L=9.4~{\rm fb}^{-1}$ ($L=0.94~{\rm fb}^{-1}$) at $R=0.03\%$, corresponding
to roughly two years of running at each $R$ for optimistic (pessimistic)
instantaneous luminosity assumptions. Efficiencies and cuts
employed appear in \cite{bbgh}. The effects of bremsstrahlung are included.}
\small
\begin{center}
\begin{tabular}{|c|c|c|c|c|}
\hline
Quantity & \multicolumn{4}{c|}{Errors} \\
\hline
\hline
{\bf Mass (GeV)} & {\bf 80} & {\bf $M_Z$} & {\bf 100} & {\bf 110} \\
\hline
$\sigma_cB(b\overline b)$ & $0.7\%(2.2\%)$ & $6.3\%(20\%)$ & $ 1.2\%(3.8\%)$ &
$ 0.9\%(2.8\%)$ \\
\hline
$\sigma_cB(WW^\star)$ & $-$ & $-$ & $ 8.2\%(26\%)$ &
$ 3.8\%(12\%)$ \\
\hline
$\sigma_cB(ZZ^\star)$ & $-$ & $-$ & $-$ &
$ 60\%(190\%)$ \\
\hline
$\Gamma_H$ & $ 6\%(19\%)$ & $ 63\%(200\%)$ & $ 14\%(45\%)$ &
$ 8\%(25\%)$ \\
\hline
\hline
{\bf Mass (GeV)} & {\bf 120} & {\bf 130} & {\bf 140} & {\bf 150} \\
\hline
$\sigma_cB(b\overline b)$ & $0.9\%(2.8\%)$ & $1.4\%(4.4\%)$ & $ 2.4\%(7.6\%)$ &
$ 6.6\%(21\%)$ \\
\hline
$\sigma_cB(WW^\star)$ & $2.4\%(7.7\%)$ & $1.8\%(5.7\%)$ & $ 1.6\%(5.0\%)$ &
$ 1.8\%(5.6\%)$ \\
\hline
$\sigma_cB(ZZ^\star)$ & $15\%(46\%)$ & $7.9\%(25\%)$ & $ 6.3\%(20\%)$ &
$ 7.0\%(22\%)$ \\
\hline
$\Gamma_H$ & $ 6.3\%(20\%)$ & $ 6\%(19\%)$ & $ 5.4\%(17\%)$ &
$ 4.7\%(18\%)$ \\
\hline
\end{tabular}
\end{center}
\label{fmcerrorsrc}
\end{table}
Let us now compare to the $r_c$-ratio technique.
We employ the same total of 4 years of operation
as considered for the three-point scan, but always with $E=M_H$.
As noted earlier, a good sharing of time is to devote two years
to running at $R=0.003\%$, accumulating $L=2~{\rm fb}^{-1}$ ($L=0.2~{\rm fb}^{-1}$)
for optimistic (pessimistic) instantaneous luminosity, and a
second two years to running at $R=0.03\%$, corresponding to [using
the luminosity scaling law of Eq.~(\ref{lform})] $L=9.4~{\rm fb}^{-1}$ ($L=0.94~{\rm fb}^{-1}$)
of accumulated luminosity. For $R=0.03\%$,
$\Gamma_H/\sigma_M^{\rm max}\sim 0.1$ and, as noted earlier, each
$\sigma_c B_F$ cross section is decreased by almost
the same factor by which the luminosity has increased, leaving
the number of signal events unchanged.
However, the background is increased by a factor of 4.7.
A complete calculation is required. This was performed in \cite{ratio}.
The various $\sigma_c B_F$ statistical errors
are summarized in Table~\ref{fmcerrorsrc} along with the
$\Delta\Gamma_H/\Gamma_H$ statistical error computed by combining all the
listed final state channels and following the procedure of
Eq.~(\ref{sigcerror}).
We observe that the ratio technique becomes superior to the scan
technique for the larger $M_H$ values ($M_H>130~{\rm GeV}$). This
is correlated with the fact that $\Gamma_H/\sigma_M^{\rm min}$ (where $\sigma_M^{\rm min}$
is that for $R=0.003\%$)
becomes substantially larger than 1 for such $M_H$. In
particular, for larger $M_H$,
$\Gamma_H/\sigma_M^{\rm central}$ is in a range such that $|s|$
and, consequently, the error in $\Gamma_H$ will be minimal.
Thus, the two techniques are actually quite complementary --- by employing
the best of the two procedures, a very reasonable determination of $\Gamma_H$
and very precise determinations of the larger channel rates
will be possible for all $M_H$ below $2M_W$.
Finally, we again note that the $r_c$-ratio technique
has the advantage that the $\Delta\sigma_M$-induced systematic error
in $\Gamma_H$ is equal to $\Delta\sigma_M/\sigma_M$ and will, therefore,
be smaller by a factor of about 2.5 (if $M_H\leq 120~{\rm GeV}$)
for the $r_c$-ratio technique than
for the scan technique and that there is no $\Delta\sigma_M$-induced
systematic error in the $B_{\ell^+\ell^-}B_F$ determinations.
Thus, although the statistical errors for the $r_c$-ratio
procedure are larger than for the scan procedure when $M_H<130~{\rm GeV}$,
if the fractional error in $\sigma_M$ is substantially larger
than $0.01$, the $r_c$-ratio could have net overall (statistical
plus systematic) error that is smaller than the scan technique
down to $M_H$ values significantly below $130~{\rm GeV}$.
\bigskip
\noindent {\bf{The lightest PNGB}} - The $s$-channel
production of the lightest neutral
pseudo-Nambu-Goldstone boson (PNGB) ($P^0$), present in models of
dynamical breaking of the electroweak symmetry which have a
chiral symmetry larger than $SU(2)\times SU(2)$, has recently been
explored \cite{mumu,mumualso}. In the broad class
of models considered in \cite{mumu}, the $P^0$ is of particular interest
because it contains only down-type techniquarks (and charged technileptons)
and thus has a mass scale that is most naturally set by the
mass of the $b$-quark. Other color-singlet PNGB's will have masses
most naturally set by $m_t$, while color non-singlet PNGB's will
generally be even heavier. The $M_{\pzero}$ mass range, that is typically
suggested by technicolor models \cite{technicolor}, is
$10~{\rm GeV}<M_{\pzero}<200~{\rm GeV}$.
Discovery of the $P^0$ in the $gg\toP^0\to\gamma\gam$
mode at the Tevatron Run II and at the
LHC will almost certainly be possible unless its
mass is either very small ($\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 30~{\rm GeV}$?) or very large ($\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}}
200~{\rm GeV}$?), where the question marks are related to uncertainties in
backgrounds in the inclusive $\gamma\gam$ channel.
Run I data at Tevatron can already be used to exclude a $P^0$ in the
$50-200~{\rm GeV}$ mass range for a number of technicolors $N_{TC}> 12-16$.
In contrast, an $e^+e^-$ collider, while able to discover the $P^0$
via $e^+e^-\to\gamP^0$, so long as $M_{\pzero}$ is not close to $M_Z$,
is unlikely (unless the TESLA $500~{\rm fb}^{-1}$ per year option
is built or $N_{TC}$ is very large) to be able to determine the rates for
individual $\gamma F$ final states ($F=b\overline b,\tau^+\tau^-,gg$
being the dominant $P^0$ decay modes) with sufficient accuracy such as
to yield more than very rough indications about the
parameters of the technicolor model. The $\gamma\gam$ collider mode of
operation at an $e^+e^-$ collider would allow one to discover and study
the $P^0$ with greater precision.
A $\mu^+\mu^-$ collider would play a very special role
with regard to determining key properties of the $P^0$.
In particular, the $P^0$, being comprised
of $D\overline D$ and $E\overline E$ techniquarks, will naturally have couplings
to the down-type quarks and charged leptons of the SM.
Thus, $s$-channel production ($\mu^+\mu^-\toP^0$) is predicted
to have a substantial rate for $\sqrt s\simM_{\pzero}$.
Because the $P^0$ has a very narrow width (see
Fig. \ref{fig6}), not much
larger than that of a SM-Higgs boson of the same mass,
in order to maximize this rate
it is important that one operates the $\mu^+\mu^-$ collider
so as to have extremely small beam energy spread, $R=0.003\%$.
For such an $R$, the resolution in $E=\sqrt s$ of the muon collider,
$\sigma_E$, is of order $\sigma_E\sim 1~{\rm MeV} (E/50~{\rm GeV})$, whereas
the $P^0$ width, $\Gamma_{\pzero}$, varies from $2~{\rm MeV}$ to $20~{\rm MeV}$ as $M_{\pzero}$
ranges from $50~{\rm GeV}$ up to $200~{\rm GeV}$ (small differences with
respect to \cite{mumu} come from running fermion masses). Thus,
$\sigma_E<\Gamma_{\pzero}$ is possible and leads to very high $P^0$
production rates for typical $\mu^+\mu^-\toP^0$ coupling strength.
Assuming that the $P^0$ is discovered at the Tevatron, the LHC
or (as might be the only possibility if $M_{\pzero}$ is very small)
at an $e^+e^-$ collider (possibly operating in the $\gamma\gam$
collider mode), the $\mu^+\mu^-$ collider could quickly (in less
than a year) scan the mass range indicated by the previous
discovery (for the expected uncertainty in the mass
determination) and center on $\sqrt s\simeqM_{\pzero}$ to within
$<\sigma_M$. A first very rough estimate of $\Gamma_{\pzero}$
would also emerge from this initial scan. One would then proceed
with a dedicated study of the $P^0$. One technique would be
to use the optimal three-point scan \cite{mumu}
of the $P^0$ resonance (with measurements
at $E=M_{\pzero}$ and $E=M_{\pzero}\pm 2\sigma_M$
using $R=0.003\%$). The three-point scan would determine with high
statistical precision all the $\mu^+\mu^-\toP^0\to F$ channel
rates and give a reasonably accurate measurement of
the total width $\Gamma_{\pzero}$. For the particular
technicolor model parameters analysed in \cite{mumu}, 4 years
of the pessimistic yearly luminosity ($L_{\rm year}=0.1~{\rm fb}^{-1}$)
devoted to the scan yields the results presented
in Fig.~19 of \cite{mumu}.~\footnote{This figure gives the errors
before taking into account the possible variation of
the luminosity with $M_{\pzero}$. If the muon collider
is built so that the $\sqrt s=100~{\rm GeV}$ luminosity is maximized,
then $L_{\rm year}$ will be smaller (larger)
than the $\sqrt s=100~{\rm GeV}$ value for smaller (larger) $\sqrt s$.
The relevant luminosity scaling laws are those given in
Eq.~(7.2) of Ref.~\cite{mumu}. The effects upon the statistical errors
quoted below of such luminosity scaling are given in
Fig.~20 of \cite{mumu}, but will not be included in our discussion here.}
Sample statistical errors for
$\sigma_cB(P^0\to {\rm all})$ and $\Gamma_{\pzero}$ taken from this figure
at $M_{\pzero}=60~{\rm GeV}$, $80~{\rm GeV}$, $M_Z$, $110~{\rm GeV}$, $150~{\rm GeV}$ and $200~{\rm GeV}$
are given in Table~\ref{fmcerrorspgb}.
\begin{table}[h]
\caption[fake]{\baselineskip 0pt Fractional
statistical errors ($1\sigma$) for
$\sigma_cB(P^0\to {\rm all})$ (combining $b\bar b$, $\tau^+\tau^-$,
$c\bar c$ and $gg$ tagged-channel rates --- see \cite{mumu})
and $\Gamma_{\pzero}$ for $s$-channel $P^0$ production at the muon collider.
We compare results for an $R=0.003\%$ three-point scan
with total integrated luminosity of $L=0.4~{\rm fb}^{-1}$ (corresponding
to four years of running at pessimistic luminosity, with
distribution $L/5$ at $E=M_{\pzero}$,
$2L/5$ at $E=M_{\pzero}+2\sigma_M$ and $2L/5$ at $E=M_{\pzero}-2\sigma_M$)
to results obtained using the $r_c$-ratio technique
assuming accumulated luminosities at $E=M_{\pzero}$
of $L=0.2f~{\rm fb}^{-1}$ at $R=0.003\%$ and
$L=0.94(2-f)~{\rm fb}^{-1}$ at $R=0.03\%$ (corresponding
to roughly $2f$ years of running at $R=0.003\%$ and $(4-2f)$
years of running at $R=0.03\%$ for pessimistic
instantaneous luminosity assumptions). $f$ (tabulated below) is chosen
to minimize the error in $\Gamma_{\pzero}$. We employ the
efficiencies, cuts and tagging procedures described in \cite{mumu}.
The effects of bremsstrahlung are included.}
\small
\begin{center}
\begin{tabular}{|c|c|c|c|c|c|c|}
\hline
Quantity & \multicolumn{6}{c|}{Errors for the scan procedure} \\
\hline
\hline
{\bf Mass (GeV)} & {\bf 60} & {\bf 80} & {\bf $M_Z$} & {\bf 110}
& {\bf 150} & {\bf 200} \\
\hline
$\sigma_cB$ & 0.0029 & 0.0054 & 0.043 & 0.0093 & 0.012 & 0.018\\
\hline
$\Gamma_{\pzero}$ & 0.014 & 0.029 & 0.25 & 0.042 & 0.052 & 0.10 \\
\hline
\hline
Quantity & \multicolumn{6}{c|}{Errors for the $r_c$-ratio procedure} \\
\hline
\hline
{\bf Mass (GeV)} & {\bf 60} & {\bf 80} & {\bf $M_Z$} & {\bf 110}
& {\bf 150} & {\bf 200} \\
\hline
$f$ & 0.8 & 0.7 & 0.6 & 0.8 & 0.9 & 1.0 \\
\hline
$\sigma_cB$ & 0.0029 & 0.0062 & 0.055 & 0.010 & 0.011 & 0.016 \\
\hline
$\Gamma_{\pzero}$ & 0.014 & 0.028 & 0.24 & 0.041 & 0.039 & 0.053 \\
\hline
\end{tabular}
\end{center}
\label{fmcerrorspgb}
\end{table}
In the analysis performed in \cite{mumu},
we did not consider the errors induced
by a systematic error in the energy spread. In Fig. \ref{fig7} we show
the $\Delta\sigma_M$-induced fractional errors for the branching ratio $B$
(where $B=B_{\ell^+\ell^-}B_F$ if we focus on a given final state
or $B=B_{\ell^+\ell^-}$ if we sum over all final states)
and for $\Gamma_{\pzero}$ predicted for the same three-point scan measurement
($E=M,~E=M\pm 2\sigma_M$, $R=0.003\%$) as employed
for the statistical error analysis summarized above.
Results are shown for the two cases of
$\Delta \sigma_M/\sigma_M=0.05$ (solid line) and
$\Delta \sigma_M/\sigma_M=0.01$ (dashed line).
For $R=0.003\%$, we find $\Delta\Gamma_{\pzero}/\Gamma_{\pzero}\sim
c_{\Gamma_{\pzero}} \Delta\sigma_M/\sigma_M$ with $c_{\Gamma_{\pzero}}\sim 1.5$
for $M_{\pzero}\sim 60~{\rm GeV}$ falling to $c_{\Gamma_{\pzero}}\sim 0.45$
for $M_{\pzero}\sim 200~{\rm GeV}$. For $M_{\pzero}\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 80~{\rm GeV}$, the result
is that the induced $\Delta\Gamma_{\pzero}/\Gamma_{\pzero}$
systematic errors are comparable to
the expected statistical errors even for $\Delta \sigma_M/\sigma_M=0.01$.
For example, at $M_{\pzero}=60~{\rm GeV}$ both the systematic error
and the statistical error are of order 1.5\%.
In the neighborhood of the $Z$ peak the errors
from the optimal three-point scan \cite{mumu} are largely
dominated by the $Z$ background and the $\Delta
\sigma_M/\sigma_M$ effect can be neglected if $\Delta\sigma_M/\sigma_M$
is not large. For $M_{\pzero}\sim 150~{\rm GeV}$ ($\sim 200~{\rm GeV}$),
$c_{\Gamma_{\pzero}}\sim 0.6$ ($0.45$), {\it i.e.}\
$\Delta\Gamma_{\pzero}/\Gamma_{\pzero}\sim 0.6\Delta\sigma_M/\sigma_M$
($0.45\Delta\sigma_M/\sigma_M$),
while the statistical $\Delta\Gamma_{\pzero}/\Gamma_{\pzero}$
error from the three-point scan \cite{mumu} would be of the
order $5\%$ ($10\%$). Thus, for $\Delta \sigma_M/\sigma_M\leq 0.01$
the systematic error would be much smaller than the statistical error.
As a result, for $M_{\pzero}\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 100~{\rm GeV}$, as far as the $\Delta\sigma_M$-induced
errors for $\Gamma_{\pzero}$ are concerned,
one could consider employing a value of $R$ larger than $0.003\%$.
As an example, $R=0.01\%$ could be chosen. Since $\Gamma_{\pzero}/\sigma_M$
is still $\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$ for this $R$ and $M_{\pzero}\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 100~{\rm GeV}$,
we can expect that systematic errors will still be under control.
The actual systematic errors
resulting from an $E=M$, $E=M\pm 2\sigma_M$ three-point
scan appear in Fig.~\ref{fig7}. For $M_{\pzero}\sim 150~{\rm GeV}$ ($\sim 200~{\rm GeV}$)
and $\Delta\sigma_M/\sigma_M=0.01$, one finds systematic error
of $\Delta\Gamma_{\pzero}/\Gamma_{\pzero}\sim 0.028$ ($0.022$), which is
indeed an acceptable level. However, as described in the previous Section,
despite the factor of 2.2 increase [see Eq.~(\ref{lform})] in yearly
luminosity achieved by increasing $R$ from $0.003\%$ to $0.01\%$,
the decrease in the signal to background ratio is very substantial
and would lead to worse statistical errors unless $M_{\pzero}>200~{\rm GeV}$
[for which $\Gamma_{\pzero}/\sigma_M(R=0.003\%)>5$]. Thus, for $M_{\pzero}<200~{\rm GeV}$
and typical model parameters as embodied in the choices of Ref.~\cite{mumu},
one should employ $R=0.003\%$ for the scan.
Let us now consider the $r_c$-ratio technique for the $P^0$. We will
compare to the scan technique using the choices $R=0.003\%$ for $\sigma_M^{\rm
min}$ and $R=0.03\%$ for $\sigma_M^{\rm max}$. This means $\sigma_M^{\rm
central}\sim 6.3~{\rm MeV}\,(M_{P^0}/100~{\rm GeV})$. From Fig.~\ref{fig6}, we then
find $\Gamma_{\pzero}/\sigma_M^{\rm central}\sim 2/3$ at $M_{\pzero}=50~{\rm GeV}$
rising slowly to 1.6 at $M_{\pzero}=200~{\rm GeV}$.
This region is that for which the slope $|s|$ (see Fig.~\ref{dlgdlr})
is smallest. Consequently, the error in $\Gamma_{\pzero}$
will be small if that for $r_c$ is.
The $\Delta\sigma_c/\sigma_c$ errors for 3 years of operation at
pessimistic instantaneous luminosity ($L=0.3~{\rm fb}^{-1}$) at $R=0.003\%$
were given in Fig.~19 of Ref.~\cite{mumu}. We rescale these
errors to $L=0.2f~{\rm fb}^{-1}$ (corresponding to $2f$ years of operation
at $R=0.003\%$), where $f$ will be chosen to minimize the error in $r_c$.
We also compute $\Delta\sigma_c/\sigma_c$ for $L=0.94(2-f)~{\rm fb}^{-1}$
devoted to $R=0.03\%$ running (corresponding to $4-2f$ years
of operation at this latter $R$).
The net $\Delta\sigma_c/\sigma_c$ is computed using Eq.~(\ref{sigcerror})
after combining all final state channels.
We also compute $\Delta r_c/r_c$ according to the Eq.~(\ref{sigcerror})
procedure. The corresponding statistical error for $\Gamma_{\pzero}$ is then
computed using the appropriate $|s|$ slope value. We then search
for the value of $f$ (see above) such that $\Delta\Gamma_{\pzero}/\Gamma_{\pzero}$
is smallest. The value of $f$ and the corresponding errors
for the combined-channel
$\sigma_c$ and for $\Gamma_{\pzero}$ error are given in Table~\ref{fmcerrorspgb}
for the same $M_{\pzero}$ values as considered for the three-point scan procedure.
For $\sigma_c$, the $r_c$-ratio procedure statistical errors
are very similar to the 4-year three-point scan statistical errors
for all $M_{\pzero}$ values considered. For $\Gamma_{\pzero}$,
the $r_c$-ratio procedure statistical errors are as good
as the tabulated 4-year three-point scan statistical errors
for $M_{\pzero}<110~{\rm GeV}$,
and become superior for larger $M_{\pzero}$ values where $\Gamma_{\pzero}/\sigma_M^{\rm
central}$ is significantly larger than unity.
As we have discussed, for the $r_c$-ratio procedure,
the fractional systematic error in $\Gamma_{\pzero}$
is equal to that in $\sigma_M$ and there is no systematic
error in $B\equiv B_{\mu^+\mu^-}B_F$.
In contrast, we have seen that for $M_{\pzero}<80~{\rm GeV}$
the systematic errors in $\Gamma_{\pzero}$ and $B$ from
the scan technique will be somewhat larger than $\Delta\sigma_M/\sigma_M$.
Taking $\Gamma_{\pzero}$ as an example, the scan systematic errors are of
order $1.5\Delta\sigma_M/\sigma_M$. As summarized earlier, this means that
the scan systematic error for $\Delta\sigma_M/\sigma_M=0.01$
is essentially the same as the
scan statistical error computed assuming pessimistic luminosity.
For optimistic luminosity the scan systematic error would be dominant.
Thus, the $r_c$-ratio procedure will actually give better overall
(systematic plus statistical) error for $\Gamma_{\pzero}$ and, especially,
$B$ than the scan procedure for low as well as high $M_{\pzero}$
values. This will become especially important if the luminosity available
is better than the pessimistic value or if $\Delta\sigma_M/\sigma_M>0.01$.
If a narrow resonance is observed in the $\gamma\gam$ final state at
the Tevatron or LHC, and if it has the weak coupling
to $ZZ$ that is typical of a pseudogoldstone boson,
then the $r_c$-ratio technique for precision measurements of its properties
at the muon collider is strongly recommended.
\bigskip
\noindent {\bf{Degenerate BESS}} -
The Degenerate BESS model \cite{DBESS} describes
two isotriplets
of nearly degenerate vector resonances $(\vec L,\vec R)$
characterized by two parameters $(M,g'')$, the common mass (when
the EW interactions are turned off) and their gauge coupling. The
main feature of the model is the decoupling property, which
implies very loose bounds from existing precision experiments
\cite{vancouver} as shown in Fig. \ref{fig8}. Another consequence is that
the decay of the resonances into pairs of ordinary gauge vector
bosons is quite depressed. The total widths for the two neutral
states $(L_3,R_3)$ are given by
\begin{equation}
\Gamma_{L_3,R_3}=M\,h_{L_3,R_3}(g/g'')\,,
\end{equation}
where $g$ is the weak coupling constant. The behaviour of the
total widths as functions of $g/g''$ is
shown in Fig. \ref{fig9}, whereas for $g/g''\ll
1$ one has
\begin{equation}
h_{L_3}\approx 0.068\left(\frac g{g''}\right)^2,~~~ h_{R_3}\approx
0.01\left(\frac g{g''}\right)^2\,.
\end{equation}
The ratio of the widths, $\Gamma_{L_3}/\Gamma_{R_3}$, in the
interval $0\le g/g''\le .5$ approximately varies between $0.15$
and $0.07$. The weak interactions break the mass degeneracy
giving rise to the mass splitting
\begin{equation}
M^s=M_{L_3}-M_{R_3}=M\,h_{M^s}(g/g'')\,.
\end{equation}
The behaviour of $h_{M^s}$ for $g/g''\ll 1$ is
\begin{equation}
h_{M^s}\approx (1-\tan^2\theta_W)\left(\frac
g{g''}\right)^2\approx 0.7 \left(\frac g{g''}\right)^2\,.
\end{equation}
The branching ratios into charged lepton pairs are almost parameter
independent and rather sizeable
\begin{equation}
B(L_3\to\ell^+\ell^-)\approx
4.5\%,~~~~B(R_3\to\ell^+\ell^-)\approx 13.6\%\,.
\end{equation}
In this Section, we will assume that $\sigma_M$ is much smaller than
$M^s$, and therefore we can apply the previous analysis to
the two resonances separately. The case $\sigma_M\approx
M^s\gg
\Gamma_{L_3,R_3}$ will be studied in Section 4.
By combining Fig. \ref{fig2} with Fig. \ref{fig9},
we easily obtain the $\Delta\sigma_M$-induced
fractional errors in the branching ratios and in
the widths of the two resonances $L_3$ and $R_3$ as functions of
$g/g''$ for different choices of $\sigma_M$ [equivalently, $R$; see Eq.
(\ref{spread})]. The results are given in Figs. \ref{fig10} and \ref{fig11} for
$L_3$ and $R_3$, respectively, for the choices
$\Delta\sigma_M/\sigma_M=1\%$ and 5\%, and for $R=1\%$
and 0.1\%. (Because the resonance widths are large,
we do not need the very small values of $R$ required
to study the SM Higgs or the $P^0$.)
Combining these results with the bounds on the
portion of parameter space still allowed
by precision experiments, one can put lower limits on the masses of
the resonances such that they have not been excluded and yet
one is able to measure the widths with no more than a
given systematic uncertainty from $\Delta\sigma_M$.
For instance, in the case of a machine with $R=1\%$ (typical of an
$e^+e^-$ collider) and for $\Delta\sigma_M/\sigma_M=1\%$,
one finds, from Fig. \ref{fig9}, that
$g/g''\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.3$ is needed to avoid $\Delta\sigma_M$-induced
errors in $\Gamma_{L_3}$ above $2\%$. From Fig. \ref{fig8}, the
portion of parameter space that is still allowed by precision
experiment can be roughly expressed by
\begin{equation}
\frac{M}{1000~{\rm GeV}}\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 2.17\left(\frac g{g''}+0.01\right)\,.
\end{equation}
Therefore, the $g/g''\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.3$ requirement converts to the requirement
that $M\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 670~{\rm GeV}$.
This means that, at lepton colliders with $R=1\%$ and
$\Delta\sigma_M/\sigma_M=1\%$, one will be able to measure
$\Gamma_{L_3}$ (for an allowed $L_3$ resonance)
with a $\Delta\sigma_M$-induced error
of less than 2\% only if the mass
of the resonance is greater than 670 GeV. Similar
considerations apply to $R_3$, where keeping the systematic
error in $\Gamma_{R_3}$ below $2\%$ requires
$g/g''\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.7$, leading to $M\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 1540$ GeV.
In the case of a muon collider, $R=0.1\%$ can be achieved
while maintaining large luminosity. In this case,
the $\Delta\sigma_M$-induced fractional error for $L_3$
is below $2\%$, for $\Delta\sigma_M/\sigma_M=1\%$, if $g/g''\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.12$,
which converts to $M\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 280~{\rm GeV}$ for $L_3$
resonances not already excluded by the precision data. The corresponding
limits for $R_3$ are $g/g''\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.3$
and $M\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 670~{\rm GeV}$. In particular, we see from Fig. \ref{fig8} that,
at a machine working at the top threshold (about 350 GeV), only
resonances with $g/g''\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.14$ have not already been
excluded by the precision data. Therefore,
only a machine with $R\le 0.1\%$ would be able to measure the
width and branching ratios of an allowed resonance without
encountering significant systematic uncertainty
coming from $\Delta\sigma_M/\sigma_M\sim 0.01$.
\bigskip
\section{Analysis for nearly degenerate resonances}
\bigskip
In this Section we will discuss nearly degenerate
resonances, {\it i.e.}\ the case of a mass splitting much smaller
than the average mass of the resonances, $M^s=M_2-M_1\ll M=(M_1+M_2)/2$.
We will be interested in the case where the energy spread
of the beam is of the same order of magnitude as the mass
splitting between the two peaks, $\sigma_M\approx M^s$. We
will also assume that the widths of the two resonances are much
smaller than the mass splitting, {\it i.e.}\ $\Gamma_1,\,\Gamma_2\ll
M^s$. It follows also that $\Gamma_1,\,\Gamma_2\ll\sigma_M$. In this
approximation, we can safely describe the cross section as the sum
of two Breit-Wigner functions, and furthermore we can use the
narrow width approximation. Therefore,~\footnote{We
focus on the total cross section, but it should be kept in mind
that we could also consider the cross section in a given final state.
In this case, $\Gamma(R_i\to\ell^+\ell^-)$ ($i=1,2$) should be replaced by
$\Gamma(R_i\to\ell^+\ell^-)B_F$ in all that follows.
The $\Delta\sigma_M$-induced systematic errors on this product
would be the same as for $\Gamma(R_i\to\ell^+\ell^-)$.}
\begin{equation}\label{bw2}
M^2\sigma_c=
B_{\ell^+\ell^-}^1\Phi(x_1,\gamma_1)+B_{\ell^+\ell^-}^2\Phi(x_2,\gamma_2)\,,
\end{equation}
where $B_{\ell^+\ell^-}^1$ and $B_{\ell^+\ell^-}^2$ are the branching
ratios of the resonances into $\ell^+\ell^-$, and
\begin{equation}\label{def2}
x_1=\frac{E-M_1}{\sigma_M},~~~x_2=\frac{E-M_1-M^s}{\sigma_M}=x_1-m^s
\end{equation}
and
\begin{equation}\label{def3}
M_2=M_1+M^s,~~~m^s=\frac{M^s}{\sigma_M}\,.
\end{equation}
We have also assumed $M\approx M_1\approx M_2$. For a Gaussian
beam we recall that the function $\Phi(x,\gamma)$ defined in Eq.
(\ref{scaled-conv}) is given by
\begin{equation}\label{Phi}
\Phi(x,\gamma)=\sqrt{\frac \pi 2} (2 j+1)\int_{-\infty}^{+\infty}e^{-(x-y)^2/2}
\frac{\gamma^2}{y^2+\gamma^2/4}dy\,.
\end{equation}
We may evaluate this expression by
performing the Fourier transforms of the Gaussian distribution and
of the Breit-Wigner and then taking the inverse Fourier
transform. We get
\begin{equation}\label{integration}
\Phi(x,\gamma)= (2j+1)\pi^{3/2}\gamma\int_{-\infty}^{+\infty}dp
e^{-p^2/2+ipx}
\left(\theta(p)e^{-p\gamma/2}+
\theta(-p)e^{+p\gamma/2}\right)\,.
\end{equation}
Since we are assuming that the resonance widths are much
smaller than the energy spread, we may evaluate Eq.~(\ref{integration})
in the $\gamma\to 0$ approximation, yielding
\begin{equation}\label{approx}
\Phi(x,\gamma)=(2j+1)\gamma\sqrt{2}\pi^{3/2} e^{-x^2/2}\,,
\end{equation}
as earlier given in Eq.~(\ref{smallgam}) for $j=0$. As previously noted,
this expression shows that when the energy spread is much bigger than
the width the convolution gives rise to a Gaussian function
with spread $\sigma_M$, and we loose any information
about the total width. In fact, the total cross section $\sigma_c$
depends only on the product
$B_{\ell^+\ell^-}\gamma=\Gamma(R\to\ell^+\ell^-)/\sigma_M$.
In the case of two resonances with $\gamma_1,\,\gamma_2\to 0$, we
thus find
\begin{equation}\label{final cross}
M^2\sigma_c=(2j+1)\sqrt{2}\pi^{3/2}\left( g_1e^{-x_1^2/2}+g_2
e^{-(x_1-m^s)^2/2}\right)\,,
\end{equation}
where
\begin{equation}
g_i=\gamma_iB^i_{\ell^+\ell^-}=\frac{\Gamma(R_i\to\ell^+\ell^-)}{\sigma_M}\,.
\end{equation}
The function (\ref{final cross}) is invariant under the substitution
\begin{equation}
g_1\leftrightarrow g_2,~~~x_1\leftrightarrow m^s-x_1
\label{invariance}
\end{equation}
The behaviour of the function is characterized by the ratio
\begin{equation}
\frac{g_2}{g_1}=\frac{\Gamma(R_2\to\ell^+\ell^-)}{\Gamma(R_1\to\ell^+\ell^-)}
\equiv
a
\end{equation}
and by $m^s$. For small $m^s$, the convolution of the Gaussian with the
two Breit-Wigners has a single maximum in between 0 and $m^s$
depending on the value of $a$. For instance, for $a=1$ the
maximum is at $m^s/2$. In this situation, the second derivative
of the function (\ref{final cross}) has two zeros corresponding
to the changes of curvature before and after the
peak. By increasing $m^s$ the second derivative acquires a third
zero (in fact, a double zero). This is due to the effect of the smaller
Breit-Wigner which gives rise to a further change of the
curvature. Just to get an idea, we list in Table 5, for several choices of
$a$, the critical value $m^s_1$ of $m^s$ at which this
third zero occurs.
As $m^s$ is increased further, there comes a point
at which the two Breit-Wigner maxima start to
show up. The minimum value of $m^s$ required
to see two maxima, $m^s_2$,
is given as a function of the ratio $a$ in Table \ref{doublezero}.
Notice that the invariance (\ref{invariance}) implies $m^s_i(a)=m^s_i(1/a)$.
\begin{table}[bht]
\caption{Values of $m^s$ at which the double
zero of the second derivative of the function (\ref{final cross})
occurs ($m^s_1(a)$) and at which the two maxima of the function
(\ref{final cross}) start to show up ($m^s_2(a)$).}
\begin{center}
\begin{tabular}{|c|c|c|}
\hline
$a$ & $m^s_1(a)$ & $m^s_2(a)$ \\
\hline\hline
1 & 2 &2\\
\hline
2 & 1.85 & 2.63 \\
\hline
3 & 2.02 & 2.85 \\
\hline
4 & 2.20 & 2.98\\
\hline
5 & 2.31 & 3.08\\
\hline
\end{tabular}
\end{center}
\label{doublezero}
\end{table}
We can now discuss the type of measurements necessary to
determine the parameters $M_1$, $M_2$, $g_1$ and $g_2$.
In this discussion, we will use the notation $x\equiv x_1$.
We first determine the overall location in energy of the
resonance structure by locating the absolute maximum
of the cross section.
For small $m^s$, such that the individual resonance peaks
are unresolved, the cross section has a single maximum near
the location of the resonance with the larger $g$; we will assume
that it is $g_2$ which is largest. For $m^s$ large enough
that the peaks are resolved, we may locate the larger maximum.
We denote the location of the maximum in the cross section by $E_{\rm max}$,
and choose this as our first energy setting. We write
\begin{equation}
E_{\rm max}=\sigma_{M} x_{\rm max} +M_1\,,
\label{energyscale}
\end{equation}
where $x_{\rm max}$ is a function of $a$ and $m^s$ which can be
evaluated numerically from Eq. (\ref{final cross}). For instance,
for $a\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 2$ it is a good approximation to assume $x_{\rm
max}\approx m^s$ (specially for $m^s$ bigger than the critical
value $m^s_2$), implying $E_{\rm max}\simeq M_2$.
The value of the cross section at $x_{\rm max}$ provides a second
input into determining $M_{1,2}$ and $g_{1,2}$.
To complete the process of determining these four parameters,
we need two more measurements. If $m^s$ is larger than $m_2^s$,
so that a second (lower) maximum is present, we can use the
location of the second maximum and the cross section value
at this second maximum as our two additional measurements.
We effectively have four equations in four unknowns, two equations
involving the derivatives of Eq.~(\ref{final cross}) and two involving
the absolute magnitude of Eq.~(\ref{final cross}).
If no second maximum is present, then we must effectively determine
the slope of $\sigma_c(x)$ of Eq.~(\ref{final cross}) at some energy
location away from $x_{\rm max}$ and determine
the cross section at this same location. Measurements of $\sigma_c(x)$
at two nearby values of $x$ away from $x_{\rm max}$ are needed.
That these approaches are really
equivalent becomes apparent when we realize that the $m^s>m_2^s$ procedure
of locating the second maximum actually requires measuring the slope
of $\sigma_c(x)$ and finding its second zero.
(Recall that this discussion assumes infinite statistics so that
we will end up effectively computing the minimum error that
will be induced by uncertainty in $\sigma_M$.)
In the absence of a second maximum, the choice of the
other two energies can be difficult if $m^s$ is smaller than the
critical value $m^s_1$. In fact, in this case the
convolution of the two Breit-Wigner looks very much like the
convolution with a single Breit-Wigner. Not surprisingly,
to obtain reasonable errors
it is necessary that $\sigma_M$ be such that $m^s$ is at least bigger than
$m^s_1$. Let us assume that for $m^s>m^s_1$ we can
approximately locate the energy corresponding to $x\sim 0$ and that we
measure the cross section at two points
in its vicinity (as well as at $x=x_{\rm max}$). If
$m^s$ is greater than $m^s_2$, we may continue to employ the
above procedure or we may use the alternative procedure
outlined earlier based on the fact that the
cross section has two maxima, one near $x=0$ and one near $x=m^s$;
in the alternative procedure, we measure the energy and
the cross section at the two peaks.
We consider first the former procedure that is the only choice if $m^s<m^s_2$.
We start our analysis by taking $a=2$ and then later discuss
the modifications for different values of $a$. From the
measurement which fixes the energy scale (see Eq. (\ref{energyscale}))
we find the following relation between the parameter errors
and the uncertainty in the energy spread:
\begin{equation}
\left(1+\frac{\Delta\sigma_M}{\sigma_M}\right)x_{\rm max}(m^s+\Delta m^s)
+\Delta m_1-x_{\max}(m^s)=0\,,
\label{energy1}
\end{equation}
where $\Delta M_1=\sigma_M\Delta m_1$ is the error in $M_1$ and
$\Delta m^s$ is the error in $m^s$. From this equation we can
eliminate $\Delta m_1$ in terms of the other
errors. This has been done by using the approximation $x_{\rm
max}\approx m^s$. The errors on $M^s$ and on the partial
widths can then be determined using the three
cross sections --- $\sigma_c(x_{\rm max})$
and $\sigma_c(x)$ at two other $x$ values near 0 ---
following a procedure analogous to that discussed for
a single resonance using Eq. (\ref{condition1}).
In particular, we assume measurements of the
cross section at $x_{\max}$ and at $x=0.1$ and $x=0.2$. The
resulting fractional errors for $\Gamma(R_i\to\ell^+\ell^-)$ ($i=1,2$)
and $M^s$ are given in Fig. \ref{fig12} as a function of $M^s/\sigma_M$
for $\Delta\sigma_M/\sigma_M=5\%$. As
expected, the errors grow rapidly once $m^s$ falls below $m^s_1$.
As $m^s$ is increased above $m^s_1$, there is a change of curvature
in the fractional error curves around the critical value $m^s_2$,
after which the fractional errors approach asymptotic limits.
Notice that the asymptotic value of $\Delta M^s/M^s$ is zero,
whereas $\Delta\Gamma(R_i\to\ell^+\ell^-)/\Gamma(R_i\to\ell^+\ell^-)\to
\Delta\sigma_M/\sigma_M$ for both $i=1,2$. This is because the cross sections
depend only on the ratio $\Gamma(R_i\to\ell^+\ell^-)/\sigma_M$.
In Fig. \ref{fig13} we represent the same quantities but for $a=3$, and we
see results very similar to those for $a=2$ except
for changes due to the different values of the critical
points $m^s_1$ and $m^s_2$. We have tried different choices for the
two measured points off the maximum, varying them up to
$x=0.3$, without any significant change in the results.
We now consider the alternative procedure outlined earlier that
is possible when two cross section maxima become visible,
that is when the mass splitting is
bigger than $m^s_2$. In this case, the four
measurements for determining $M_{1,2}$ and $g_{1,2}$
are the energy locations of the two maxima and the
cross sections at these two maxima. The measurement of the energy of the
second maximum gives rise to a condition similar to the one of Eq.
(\ref{energy1}). The resulting fractional errors
in $M^s$ and $\Gamma(R_i\to\ell^+\ell^-)$ are given in Fig. \ref{fig14}
(for $a=2$), and are essentially the same as obtained
in the previous procedure when $m^s>m^s_2$.
The basic conclusion from these analyses is that for $m^s$ of
the order of the critical value $m^s_2$, the fractional
errors in the parameters of the resonances are of the order of
the fractional error in $\sigma_M$. For smaller $m^s$ the errors
become very large. For $m^s$ significantly bigger
than $m^s_2$, the fractional error in the
mass splitting rapidly approaches zero while the fractional errors for
the $\Gamma(R_i\to\ell^+\ell^-)$ partial
widths become equal to the fractional error in the energy spread.
In short, we can use the critical value $m^s_2$
in order to discriminate between a good and a bad determination
of the mass splitting.
\bigskip
\section{Application to Degenerate BESS}
\bigskip
In Degenerate BESS one can show
that the condition $M^s
\gg\Gamma_L,\,\Gamma_R$ is rather well satisfied (by one and two
orders of magnitude respectively). Therefore we can apply the
analysis of the previous Section. From Fig. \ref{fig15}, we see that the
value of $a$ is almost constant and approximately 2.2 for $g/g''$
up to 0.2, and then $a$ increases up to $\approx 4$ for $g/g''=0.5$.
As discussed in the last Section, we can use the values of $m^s_2$
given in Table \ref{doublezero} in order to determine
the minimum value of $M^s/\sigma_M$ needed
in order to make a good determination of the mass
splitting. For $a=2.2$ one finds that the minimum
value is $m^s_2=2.68$. As in our earlier single resonance discussions,
for any fixed value of the energy resolution $R$, we
convert this bound into lower bounds for $g/g''$ and for the mass
$M$ of nearly degenerate resonances that have not already been
excluded by precision experimental data. From
$M^s/\sigma_M\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 2.68$ we get
\begin{eqnarray}
R=1\% &\to& \frac g{g''}\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.16,~~~ M\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 370~{\rm GeV}\,,\nonumber\\
R=0.1\% &\to& \frac g{g''}\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.05,~~~ M\mathrel{\raise.3ex\hbox{$>$\kern-.75em\lower1ex\hbox{$\sim$}}} 130~{\rm GeV}\,.\nonumber
\end{eqnarray}
We see that for $R=1\%$ a machine with energy near the top threshold would
just be at the border of being able to accurately measure
the mass difference between the $L_3$ and $R_3$ resonances not already
excluded by precision data.
\bigskip
\section{Conclusions}
We have considered the production of a narrow resonance
via $s$-channel collisions of leptons ($\ell=e$ or $\mu$).
Here, a `narrow' resonance is defined as one that has
width $\Gamma$ substantially smaller than the beam energy spread $\Delta E_{\rm
beam}$ that is natural for the collider (and therefore is associated with
the largest instantaneous luminosity). It will be convenient to use
the parameterization $\Delta E_{\rm beam}=0.01R\,E_{\rm beam}$,
where $R$ is in per cent. For example, at a muon collider
with center of mass energy $E\sim 100~{\rm GeV}$,
$R\sim 0.12\%$ allows for maximal $L$
and $L$ declines rapidly as $R$ is forced
to smaller values by compression techniques. A resonance with width
$\Gamma\ll 0.001M$ would then be narrow. Our focus has been on the
systematic error that might be introduced into measurements of the
parameters of a narrow resonance due to systematic uncertainty in the value of
$\Delta E_{\rm beam}/E_{\rm beam}$.
The important parameters that can be measured are the branching
ratio of the resonance
into the charged leptons ({\it i.e.}\ those that are being collided), the product
of this leptonic branching ratio times that for
the resonance to decay to a particular final state,
and the total width of the resonance. We examined four methods
for determining the resonance parameters: (1) a scan of the resonance;
(2) sitting on the resonance and changing the beam energy resolution;
(3) measurement of the cross section in the $\ell^+\ell^-$
final state; and (4) measurement of the Breit-Wigner area.
Methods (3) and (4) avoid the introduction of systematic errors
due to uncertainty in $\Delta E_{\rm beam}/E_{\rm beam}$,
but for the integrated luminosities that are anticipated to be available
the statistical errors associated with these techniques
would be quite large for a narrow resonance. Methods
(1) and (2) can provide resonance parameter determinations with small
statistical error. However, even in the limit of infinite statistical accuracy,
determinations of the resonance parameters
are sensitive to systematic uncertainties in $\Delta E_{\rm beam}$
if $\Gamma$ is not much larger than $\Delta E_{\rm beam}$.
At a muon collider, the smallest $R$ that can be achieved is
expected to be $R=0.003\%$, for which $\Delta E_{\rm beam}\sim \Gamma_H$
for a light SM Higgs boson and $\Delta E_{\rm beam}\sim \Gamma_{\pzero}/2$
for the lightest pseudo goldstone boson of a technicolor model.
Consequently, in these and other similar cases, a detailed assessment of
the systematic errors in resonance parameter determinations
introduced by uncertainty in $\Delta E_{\rm beam}$ is very important.
We have performed a general analysis to determine the (systematic) errors
in the measured resonance parameters induced by a systematic uncertainty
in $\Delta E_{\rm beam}$. We find that the
induced fractional errors in the leptonic branching ratio
(and also the product of the leptonic branching
ratio times the branching ratio into any given final state)
and in the total width can be expressed as universal
functions of the ratio $\Gamma/\sigma_M$, where $\sigma_M$
is the spread in total center of mass energy resulting
from the beam energy spreads: $\sigma_M/M=0.01R/\sqrt 2$.
In the case of the minimal three-point scan, with sampling
at $E=M,E=M\pm k\Gamma$, the error functions also depend on $k$.
For a minimal three-point scan with
$k=1$, the induced $\Delta\Gamma/\Gamma$ fractional systematic
errors were parameterized as a function of $\Gamma/\sigma_M$
in Eq.~(\ref{dgamform}). Very roughly,
for $\Gamma/\sigma_M\sim 2.5$ we find that the
induced fractional errors in $\Gamma$ and $B=B_{\ell^+\ell^-}$
or $B_{\ell^+\ell^-}B_F$ ($F$=final state) are of the order of the
fractional uncertainty $\Delta\sigma_M/\sigma_M$.
As $\Gamma/\sigma_M$ increases above 2, the fractional errors smoothly
decrease. For values of $\Gamma/\sigma_M$ below 1, the
fractional errors in the resonance parameters increase very rapidly.
For example, for $\Gamma/\sigma_M\sim 1$ ($\sim 0.2$)
the $\Delta\sigma_M$-induced
fractional systematic errors in the resonance parameters increase to
$\sim 3.5\div 4\Delta\sigma_M/\sigma_M$
($\sim 20\div 25\Delta\sigma_M/\sigma_M$),
for the $k=1$ scan. Thus, to avoid large systematic errors from
$\Delta\sigma_M$, it is imperative to operate the collider with
$\Gamma/\sigma_M$ no smaller than 1. If $R$ can be adjusted to achieve
values significantly larger than 1, one can consider how to optimize
the choice of $R$ so as to minimize the total statistical plus
systematic error. A discussion was presented leading to the
following two basic conclusions.
(a)
For a broad resonance, defined as one
with $\Gamma\gg0.001M$, one should operate the muon collider
at its natural $R$ value of order $0.12\%$. The $\Delta\sigma_M$-induced
errors will be very tiny, both because $\Gamma/\sigma_M$ is very large
and because $\Delta\sigma_M/\sigma_M$ should be small for such $R$.
(b)
For a resonance with $\Gamma<0.001M$, one will typically wish to operate
at an $R$ that is significantly
smaller than that value which would yield equal statistical and systematic
errors. In a typical case, this would mean a value of $R$ such
that $\Gamma/\sigma_M$ is larger than $5\div 10$.
Of course, if the resonance is extremely narrow,
it may happen that $\Gamma/\sigma_M$ is of order, or not much larger than, unity
even for $R=0.003\%$. In this case, it will normally
be essential to run with $R=0.003\%$ even though this
$R$ yields the smallest machine luminosity. Larger values of $R$ lead
to a drastic decline in the signal to background ratio in a typical final
state that, in turn, leads to very poor statistical errors (given
the rather slow compensating increase with $R$ of the instantaneous luminosity).
For a narrow resonance
with $\Gamma\ll 0.001M$, the technique in which one sits on the resonance peak
and measures the cross section for two different values of
$\sigma_M$ ($\sigma_M^{\rm max}>\Gamma$ and $\sigma_M^{\rm min}<\Gamma$)
is a strong competitor to the scan technique. $\Gamma/\sigma_M^{\rm central}$
(where $\sigma_M^{\rm central}=\sqrt{\sigma_M^{\rm max}\sigma_M^{\rm min}}$)
is determined by the ratio $r_c=\sigma_c(\sigma_M^{\rm
min})/\sigma_c(\sigma_M^{\rm max})$, where $\sigma_c$ is the
measured cross section.
For $\sigma_M^{\rm max}/\sigma_M^{\rm min}$ of order 5 to 20,
the statistical error in $\Gamma/\sigma_M^{\rm central}$ is smallest
if $\Gamma/\sigma_M^{\rm central}\sim 2\div 3$. The larger
$\sigma_M^{\rm max}/\sigma_M^{\rm min}$, the smaller the
statistical error for $\Gamma/\sigma_M^{\rm central}$.
For a typical choice of $\sigma_M^{\rm max}/\sigma_M^{\rm min}=10$,
one finds a statistical error of $\Delta\Gamma/\Gamma\sim
1.8\Delta r_c/r_c$ for $\Gamma/\sigma_M^{\rm central}\sim 2\div 3$.
This technique has the advantage that the
$\Delta\sigma_M$-induced systematic error in $\Gamma$ is simply
given by $\Delta\Gamma/\Gamma=\Delta\sigma_M/\sigma_M$, while there
is no systematic error in the determination of any of the
$B\equiv B_{\mu^+\mu^-}B_F$
branching ratio products ($F$ = a particular final state).
Of course, if the resonance
is very narrow ({\it e.g.}\ as narrow as a light SM Higgs boson
or a light pseudogoldstone boson), $\Gamma/\sigma_M^{\rm central}\sim 2\div 3$
will not be achievable. In this case, the best that one can do is
to employ $\sigma_M^{\rm min}$ ($\sigma_M^{\rm max}$)
as given by $R=0.003\%$ ($R=0.03\%$). The statistical error in $\Gamma$
for such a situation is typically still very good.
Let us now summarize how these results apply in the specific
cases we explored.
In the case of a three-point scan
of the SM Higgs boson, we have shown that in the region
$M_H\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 110~{\rm GeV}$ it is mandatory to have
$R\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.003\%$. In fact, even for
this very small $R$ value, $\Gamma_H/\sigma_M$ is still $\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 1$,
the very minimum needed for accurate measurements of resonance parameters.
For $\Gamma_H/\sigma_M\sim 1$ the fractional
systematic errors induced in $\Gamma_H$
from uncertainty in the beam energy spread are of order
$3\Delta\sigma_M/\sigma_M$
for a $k=2$ scan. This should be compared
to the typically-expected statistical errors tabulated
in Table~\ref{fmcerrors}. For example, for $M_H=110~{\rm GeV}$
the statistical error in $\Delta\Gamma_H/\Gamma_H$ is $\sim 5\%$
for optimistic 4-year integrated luminosity of $L=4~{\rm fb}^{-1}$ at $R=0.003\%$;
$\Delta\sigma_M/\sigma_M\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.01$ would be needed for the systematic
error to be smaller than the statistical error.
For the pessimistic 4-year integrated luminosity of $L=0.4~{\rm fb}^{-1}$, the
statistical error would be much larger ({\it e.g.}\ $\Delta\Gamma_H/\Gamma_H\sim 16\%$
at $M_H=110~{\rm GeV}$) and the $\Delta\sigma_M$-induced error would
be much smaller than the statistical error if $\Delta\sigma_M/\sigma_M\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.01$.
Note, however, that increasing $R$ is not appropriate
as this would push one into the $\Gamma_H/\sigma_M<1$ region,
implying large statistical errors and still larger systematic
errors.
For the $\sigma_M^{\rm max,min}$ on-peak ratio technique, one must
choose $\sigma_M^{\rm min}$ corresponding to $R=0.003\%$
($\Gamma_H/\sigma_M^{\rm min}\sim 1$).
Results for statistical errors were presented in Table~\ref{fmcerrorsrc}.
As a point of comparison, for optimistic (pessimistic)
instantaneous luminosity and 4 years of operation, the net production rate
error after summing over important channels is
of order 0.8\% (2.7\%) for $M_H=110~{\rm GeV}$, and the
$\Delta\Gamma_H/\Gamma_H$ statistical error
is of order 8\% (25\%). Although the statistical $\Delta\sigma_c/\sigma_c$
fractional error is somewhat smaller for the ratio technique than
for the scan technique, the $\Delta\Gamma_H/\Gamma_H$ statistical error is larger.
However, the ratio technique might still be better if $\Delta\sigma_M/\sigma_M$
were as large as $5\%$, especially
if the optimistic luminosity level is available.
This is because the systematic error in $\Gamma_H$ is equal
to $\Delta\sigma_M/\sigma_M$ for the ratio technique as opposed to
$3\Delta\sigma_M/\sigma_M$ for the scan technique.
The $r_c$-ratio technique becomes increasingly superior as the assumed
luminosity increases. For $M_H\geq 130~{\rm GeV}$, the ratio technique gives
smaller statistical errors for both $\sigma_cB$ and $\Gamma_H$
than does the scan technique (for which statistical
errors rapidly become very large). Indeed, the two procedures are nicely
complementary in that at least one of them will allow a measurement
of $\Gamma_H$ with statistical accuracy below 6\% (20\%) for optimistic
(pessimistic) luminosity.
For the lightest PNGB ($P^0$) of an extended technicolor model, the
$\Delta\sigma_M$-induced errors for the three-point
scan method can be kept smaller than in the case
of the SM Higgs boson. This is because, for typical model
parameters, the $P^0$ has a width that is larger
than that of a SM Higgs boson; $\Gamma_{P^0}/\sigma_M>2$
is quite likely for $R\approx 0.003\%$. For example,
the parameter choices of \cite{mumu} give
$\Gamma_{P^0}\sim 5~{\rm MeV}$ vs. $\sigma_M\sim 2~{\rm MeV}$ at $M_{P^0}=100~{\rm GeV}$.
As we have seen, the resulting $\Gamma_{P^0}/\sigma_M\sim 2.5$ yields
$\Delta\sigma_M$-induced resonance parameter fractional errors of order
$\Delta\sigma_M/\sigma_M$. This can be compared to the statistical errors
given in Table~\ref{fmcerrorspgb}, computed assuming the pessimistic
4-year integrated luminosity of $L=0.4~{\rm fb}^{-1}$.
For example, at $M_{\pzero}=110~{\rm GeV}$
the fractional statistical error for $\Gamma_{\pzero}$ would be $\sim 0.04$,
which is much larger than the systematic error
if $\Delta\sigma_M/\sigma_M\sim 0.01$.
For $M_{\pzero} <80~{\rm GeV}$, the statistical error in $\Gamma_{\pzero}$ declines to the
$1\%\div 3\%$ level while the systematic error for
$\Delta\sigma_M/\sigma_M=0.01$ rises to about this same level.
As $M_{\pzero}$ increases from $150~{\rm GeV}$ to $200~{\rm GeV}$, the statistical
error for $\Gamma_{\pzero}$ rises from $\sim 5\%$ to $\sim 10\%$ while
the systematic error is below $1\%$ if $\Delta\sigma_M/\sigma_M=0.01$.
For optimistic integrated luminosity of $L=4~{\rm fb}^{-1}$, the statistical
errors would be smaller than quoted above. For $\Delta\sigma_M/\sigma_M=0.01$,
the induced systematic errors would generally dominate for $M_{\pzero}\leq
80~{\rm GeV}$.
The $P^0$ resonance is sufficiently narrow that one should also
consider using the $\sigma_M^{\rm max,min}$ on-peak ratio technique
to determine $\Gamma_{\pzero}$.
For 4-year pessimistic luminosity operation we find the statistical errors
for $\Gamma_{\pzero}$ given in Table~\ref{fmcerrorspgb}.
For the on-peak ratio technique, the systematic error in $\Gamma_{\pzero}$
is equal to $\Delta\sigma_M/\sigma_M$ and for lower $M_{\pzero}$ values
would only be smaller than the statistical $\Gamma_{\pzero}$ error
if $\Delta\sigma_M/\sigma_M\mathrel{\raise.3ex\hbox{$<$\kern-.75em\lower1ex\hbox{$\sim$}}} 0.01$.
For optimistic luminosity, the $1\%$ systematic error
induced in $\Gamma_{\pzero}$ for $\Delta\sigma_M/\sigma_M=0.01$
would dominate over the statistical error for all but
$M_{\pzero}\sim M_Z$ and $M_{\pzero}>150~{\rm GeV}$.
Most importantly, the statistical and systematic errors
of the ratio technique are at least as good as, and often
better than, obtained using the scan technique.
For $M_{\pzero}\leq 110~{\rm GeV}$, the statistical $\Gamma_{\pzero}$ errors
of the ratio technique are almost the same as obtained
via the three-point scan (performed with $R=0.003\%$),
while the systemic $\Gamma_{\pzero}$ errors from the ratio technique
are smaller ($\Delta\sigma_M/\sigma_M$ vs. $\sim 1.5\Delta\sigma_M/\sigma_M$).
For $M_{\pzero}\geq 120~{\rm GeV}$, the $\Delta\sigma_M$-induced systematic
errors are comparable for the two techniques, but the statistical
errors for the ratio technique are substantially smaller than
for the three-point scan. Precision measurements
of the properties of a $P^0$ resonance would, thus, always be best
performed using the ratio procedure.
In the case of the two resonances of the Degenerate BESS
model, we have determined (for several typical $R$ values) the region
of the model parameter space for which the
fractional errors in the resonance properties ($\Gamma$, $\ldots$)
induced by $\Delta\sigma_M$
are less than a given fixed value. Induced errors are
small for large resonance masses. But, for any given choice of $R$,
as the resonance mass is decreased, while maintaining model parameter
choices such that the model is still consistent with precision experimental
data, there comes a point at which, even for
$\Delta\sigma_M/\sigma_M\sim 0.01$, the induced error
becomes large. For allowed model parameter choices yielding
a resonance mass below this, the resonance's properties
cannot be measured accurately. The lowest masses
of the resonances that correspond to
precision-data-allowed BESS model parameters,
and for which $\Delta\sigma_M$-induced
errors in the measured resonance properties are small, decrease rapidly
with decreasing $R$. As a result, the ability to achieve $R<0.1\%$
at a muon collider would be crucial for exploring the low-resonance-mass
portions of Degenerate BESS parameter space not currently excluded
by precision data.
Finally, we have performed the analysis of two nearly degenerate
resonances, a situation encountered in a number of theoretical examples,
including the Degenerate BESS model and the minimal supersymmetric model.
We focused on the case in which the
total widths of the resonances are much smaller than the mass
splitting. We have shown that, in general, the
$\Delta\sigma_M$-induced fractional error in
the measured mass splitting, $M^s$, and in the leptonic
partial widths of the resonances
(or leptonic partial widths times final state branching fraction)
depend only on $M^s/\sigma_M$ and on the ratio of the two
partial widths, $a$. The main result is that the errors are
generally big for $M^s/\sigma_M$ less than a certain critical value
(typically in the $2\div 3$ range) that is a function of $a$.
As $M^s/\sigma_M$ increases beyond
the critical value, the $\Delta\sigma_M$-induced fractional error in the
mass splitting approaches zero rapidly, whereas the fractional errors in the
partial widths approach $\Delta\sigma_M/\sigma_M$. As a
concrete case, we have discussed the application to the
two spin-1 resonances of the Degenerate
BESS model for beam energy spread of the same order as the mass splitting
between them. We determined
the regions of the model parameter space in which $M^s$ could
be measured with $\Delta\sigma_M$-induced fractional error below a given
fixed value assuming a given value of $R$. The smaller the value
of $R$ that can be used while maintaining sufficient luminosity
for small statistical errors,
the larger the fraction of allowed parameter space for which $M^s$ can
be measured with small $\Delta\sigma_M$-induced error. In particular,
$R<0.1\%$ is required if we are to be able to
separate the degenerate resonance peaks for model parameter
choices not already excluded by precision experimental data
in which the resonance masses are as low as $100~{\rm GeV}$.
\bigskip
\noindent{\bf Acknowledgements}
We wish to thank A. Blondel for the interesting discussions
during the CERN Workshop on muon colliders, where this paper was
conceived. JFG is supported in part by
the U.S. Department of Energy and by the Davis Institute for High
Energy Physics.
|
\section{How many {\it eB} neutrinos are needed?}
SK has recently presented a measurement
of the energy spectrum of recoil electrons from solar
neutrino scattering, corresponding to 504 days of data taking \cite{sk98}.
By assuming the SSM estimate of the {\it hep} neutrino flux \cite{bp98},
$\Phi_{hep}^{SSM}\simeq2\times10^{3}~\rm{cm}^{-2}~\rm{s}^{-1}$ and an undeformed
$^{8}B$ neutrino spectrum, with an arbitrary normalization,
they obtained a $\chi^2/D.O.F.=25.3/15$, corresponding to
a 4.6 \% confidence level \cite{sk98}. The poor fit
is due mainly to the behaviour of the energy-bins above $13~\rm{MeV}$.
Escribano et al. \cite{esc} suggested that a {\it hep} flux significantly
larger than the SSM estimate could reproduce the observed spectrum.
Bahcall et al. \cite{hep} have shown that a flux $\Phi_{hep}
\ge 20 \times \Phi_{hep}^{SSM}$ could
actually mimic the SK spectrum.
Alternatively, one can keep
the SSM prediction for {\it hep} neutrinos and look for other
high energy neutrino sources. Since the average energy of {\it eB} neutrinos
is roughly twice than that of {\it hep} neutrinos and since the
$\nu-e$ scattering cross section
increase linearly with energy, one expects that a flux
$\Phi_{eB}\simeq 10\times\Phi_{hep}^{SSM}\simeq 2\times 10^{4}~\rm{cm}^{-2}
~\rm{s}^{-1}$ could be sufficient to account for the high energy behaviour
of SK data.
In order to provide a quantitative estimate, let us analyse the data
by using as free parameters $\alpha=\Phi_{B}/\Phi_{B}^{SSM}$ and
$\delta=\Phi_{eB}/\Phi_{B}^{SSM} $, where $\Phi_{B}^{SSM}=5.15
\times10^{6}~\rm{cm}^{-2}~\rm{s}^{-1}$ is the SSM prediction for
the $^{8}B$ neutrino flux \cite{bp98}. We define,
in analogy with \cite{sk98}, the following $\chi^{2}$:
\begin{equation}
\chi^{2}=\sum_{i=1}^{16}
\left\{\frac{\frac{R_{i}}{SSM_{i}}
-\frac{\alpha+\delta\times B_{i}/SSM_{i}}
{(1+\delta_{i,exp}\times\beta)(1+\delta_{i,cal}\times\gamma)}}
{\sigma_{i}} \right\} ^{2}
+\gamma^{2}+\beta^{2}\ .
\label{chi2}
\end{equation}
In the previous relation $R_{i}$ is the number of
solar neutrino events observed in the i-th energy-bin;
$SSM_{i}$ \footnote{The quantities $SSM_{i}$ and $B_{i}$ have been calculated
taking into account the energy resolution of SK \cite{skres},
as described e.g. in \cite{noi}}
is the number of events
in the same energy bin due to $^{8}B$ neutrinos, for a total
flux $\Phi_{B}^{SSM}$;
$B_{i}$ is the same number due to {\it eB} neutrinos, again for a total
flux $\Phi_{B}^{SSM}$;
the quantities $\delta_{i,exp}$, $\delta_{i,cal}$, $\sigma_{i}$,
defined as in \cite{sk98}, take into account correlated
and uncorrelated theoretical and experimental errors;
the free parameters $\beta$ and $\gamma$ are used for
constraining the variation of correlated systematic errors.
For each value of $\delta$ we determined the parameters
$\alpha$, $\beta$ and $\gamma$ so as to determine the minimum of
eq. (\ref{chi2}), $\chi^{2}_{m}$, see fig. 2.
As expected, a large {\it eB} neutrino flux produces a steep increase
in the high energy tail of the Superkamiokande normalized spectrum, see fig.3.
The best-fit is obtained when $\Phi_{eB}=1.1\times10^{4}
\rm{cm}^{-2}\rm{s}^{-1}$,
corresponding to $\chi^{2}_{min}/D.O.F.=15.7/14$.
Acceptable fits are anyhow obtained for $\Phi_{eB}$ in the range
$(0.3-2)\times 10^{4} \rm{cm}^{-2}\rm{s}^{-1}$, see fig. 2.
\section{Theoretical evaluation of the {\it eB} neutrino flux}
Boron ($^8$B) is produced in the sun,
according to the following reaction
$$ ^7 Be + p \rightarrow ^8B + \gamma$$
and it undergoes $\beta^+$ decay
\begin{equation}
^8B \rightarrow ^8 Be^* + e^+ +\nu_e \rightarrow 2\alpha + e^+ +\nu_e
\label{decay}
\end{equation}
or electron capture reaction
\begin{equation}
^8B + e^-\rightarrow ^8 Be^* + \nu_e \rightarrow 2\alpha + \nu_e ~.
\label{EC}
\end{equation}
The process under consideration is an allowed transition: in fact (see
ref.\cite{Led78}) $ J^P (^8B) = J^P (^8Be^*) = 2^+$.
In this case, the ratio $R$ between electron capture probability
($\Gamma_{eB}$)
and $\beta^+ $ decay probability ($\Gamma_{\beta^+}$)
does not depend on the matrix elements of the transition operator between the
nuclear states.
A simple phase--space calculation, assuming that the electron
number density at nuclear site $n_{e}(0)$ can be approximated by
the average electron number density $n_{e}$, gives immediately
\begin{equation}
R=
\frac{1}{8 \pi}\left(\frac{hc}{m_e c^{2}}\right)^{3}
\times\left(\frac{E_{eB}}{m_{e}c^{2}}\right)^{2}\times
f^{-1}\times{n}_{e} ~,
\label{ratio}
\end{equation}
where, for later convenience, we show
explicity the dimensionless phase--space factor
associated to $\beta^{+}$ decay,
$f\simeq[(E_{eB}-m_{e}c^{2})/m_{e}c^{2}]^{5}/30\simeq7.1\times10^{5}$.
For $n_{e}\simeq5.4\times10^{25}\rm{cm}^{-3}$ as suggested by SSM,
one has $R\simeq4\times10^{-8}$ and consequently
\begin{equation}
\Phi_{eB}=R\times\Phi_{B}^{SSM}=2\times10^{-1}~\rm{cm}^{-2}~\rm{s}^{-1} ~,
\end{equation}
i.e. five orders of magnitude lower than that required to account for SK data.
It is anyhow useful to estimate $\Phi_{eB}$ with a better accuracy.
With respect to the naive estimate given previously, one
should consider the effects of interactions with the solar plasma.
The distortion of the positron wave function in the
$\beta^{+}$ decay rate can be described as a modification of
the dimensionless phase--space factor $f$, which is now given by
$f=5.70\times 10^{5}$ \cite{ec0,ec1}.
Moreover
the electron density at nucleus $n_{e}(0)$ is larger than $n_{e}$
and, consequently, the ratio $R$ has to be enhanced, with
respect to eq. (\ref{ratio}), by a factor
\begin{equation}
\omega=\frac{n_{e}(0)}{n_{e}}
\label{omega}
\end{equation}
For a precise estimate of $\omega$
one has to take into account:
{\it i)} distortion of electron wave functions in the Coulomb
field of nucleus\cite{bah62},
{\it ii)} electron capture from bound states \cite{iks},
{\it iii)} screening effects \cite{iks,bm}.
Let us discuss the problem in some detail, following the
lines of Gruzinov and Bahcall who recently produced a clear and comprehensive
analysis of the $^{7}Be$ electron capture in the sun \cite{gb}:
\begin{enumerate}
\item[\it i)]
Because of the Couloumb field of the nucleus, the wave functions of
continuum electron states differ from plane waves. The rate of electron capture
from continuum has then to be corrected by an enhancement
factor $\omega_{c}$ \cite{bah62}:
\begin{equation}
\omega_{c} = <\left|\frac{\psi_{coul}(0)}{\psi_{free}(0)}\right|^{2}>=
\left(\frac{m_{e}c^2}{kT}\right)^{\frac{1}{2}}\times
\left(Z\alpha \right)\times 2 \times \left(2\pi\right)^{\frac{1}{2}}
\times I(\beta) ~,
\end{equation}
where the average is taken over electron thermal distribution.
In the previous relation $T$ is the Sun temperature, while $I(\beta)$
is a correction factor of order unity, defined e.g. in \cite{ec2}.
For $R/R_{\odot}\simeq0.05$, which corresponds to the solar region where
the production of $^{8}B$ neutrinos is maximal,
the density enhancement at nucleus due to electron in continuum states is
$\omega_{c}=3.82$.
\item[\it ii)]
As pointed out by Iben, Kalata e Schwartz \cite{iks},
under solar conditions bound electrons give
a substantial contribution to the electron density at the nucleus.
The bound state enhancement factor is given by \cite{gb}:
\begin{equation}
\omega_{b} = \pi^{\frac{1}{2}}\times\hbar^{3}\times
\left(\frac{m_{e}kT}{2}\right)^{-\frac{3}{2}}
\sum_{n} \left(\frac{Z}{a_{0}n}\right)^{3}
\exp\left(Z^{2}e^{2}/2n^{2}a_{0}kT\right) ~,
\end{equation}
where $a_{0}$ is the Bohr radius. For $R/R_{\odot}\simeq0.05$, the
bound state enhancement factor is $\omega_{b}=2.94$. The total density
enhancement factor is then $\omega_{c}+\omega_{b}=6.76$
\item[\it iii)]
Screening effects reduce the electron density at
nucleus for both bound \cite{iks} and continuum electron states \cite{bm}.
If the temperature is
sufficiently high and if the screened potential
can by described by
\begin{equation}
V(r)=-\frac{Ze^{2}}{r}\exp(-r/R_{D}) ~,
\label{salpeter}
\end{equation}
where $R_{D}$ is the Debye radius,
by using a thermodynamical argument one finds \cite{gb}:
\begin{equation}
\omega=\exp(-Ze^{2}/kTR_{D})\times(\omega_{b}+\omega_{c}) ~.
\label{screen}
\end{equation}
For $R/R_{\odot}\simeq0.05$ the total density enhancement factor
, due to screening effects, is reduced to $\omega=5.34$.
The small difference between this value of $\omega$
and that given by \cite{gb} is due to the fact that they
have been calculated for slightly different solar regions.
Relation (\ref{screen}) is not so straightforward, especially
because of the possible inadequacies
of the Debye screening theory \cite{joh} and because
of the relatively large thermal fluctuations which could results
from the small number of ions in a Debye sphere \cite{sal}.
For the similar case of $^{7}Be$ electron capture,
Gruzinov \& Bahcall have performed a detailed analisys
of the problem, concluding that relation (\ref{omega}) is
accurate at the 2\% level.
\end{enumerate}
By using the previous relations we can determine the ratio between
electron capture and $\beta^{+}$ decay rates.
We obtain:
\begin{equation}
R=2.6\times 10^{-7}\times(1\pm0.02)
\end{equation}
This value is about 30\% larger than previous estimates \cite{ec1}
which took into account only continuum electron states contribution.
By using the SSM estimate of the $^{8}B$ neutrino flux,
which is uncertain by about 17\% \cite{bp98}, one concludes
\begin{equation}
\Phi_{eB}=R\times\Phi_{B}^{SSM}=1.3\times(1\pm0.17)~\rm{cm}^{-2}\rm{s}^{-1} ~.
\end{equation}
The predicted neutrino flux is
lower by a factor $10^{4}$ than required
to account for SK data and the calculation method is robust.
We conclude that {\it eB}
neutrinos cannot explain the spectral distributions of solar
neutrino events reported by SK.
The author thanks
M.R. Quaglia, G. Fiorentini, P. Pizzochero and P. Bortignon for
useful discussions and for earlier
collaboration on the subject of this paper.
|
\section{Introduction}
Exact Boson or Fermion solutions of the quantum N-body problem in
which every particle interacts with every other in three dimensions
are very rare. They are almost as rare in classical mechanics
although Newton solved one in Principia (1687) (see Cajori 1934 \&
Chandrasekhar 1995) and there are also some very special solutions
such as Laplace's in which the three unequal masses describe ellipses
about their centre of mass while at each time they make an {\em
equilateral} triangle. However Newton's solution was for all initial
conditions when the force on body $I$ due to body $J$ was of the form
$F_{IJ} = k m_I m_J ({\bf x}_J - {\bf x}_I)$. Newton reduced this
problem to that of $N$ harmonic oscillators relative to the centre of
mass. The quantum solution is similar to the $N$ oscillator solution
for solid state physics. The potential energy of Newton's system is
$$V = {\textstyle {1 \over 2}} \! \sum_{\ \ I \ <} \! \sum_{\! \! \! J} k m_I m_J ({\bf x} _I -
{\bf x} _J)^2 = {\textstyle {1 \over 2}} kM \Sigma m_I ({\bf x}_I -
{\overline {\bf x}})^2 \equiv {\textstyle {1 \over 2}} kM^2r^2\ . $$
Here we show that this solution may be generalised to systems in which
the total potential energy $V$ is any function of $r$. We have
already explored these systems and their generalisations in Classical
Mechanics (Lynden-Bell \& Lynden-Bell 1999). Except for Newton's
harmonic case all these systems give many-body forces in which the
force between any two bodies is approximately linear for separations
much less than the mean but with a coefficient that depends on that
current mean. Despite the strange global nature of these force laws
they may be the only non-trivial quantum many-body problems that have
been solved exactly in 3 dimensions. Only in Newton's harmonic case
do the forces reduce to simple pair-wise interactions. It could be
argued that such global forces are unnatural, however, in some
respects the resulting behaviour mimics that found in Nature. Ruth Lynden-Bell
(1995, 1996) showed that these systems can be used to give a simple
model of a phase transition that can be calculated even when $N$ is
small. Also as we now show, such forces can be used to mimic some aspects of gravitation.
For a homogeneous sphere that may pulsate in radius, $a$, the
gravitational potential energy is $V = -{3 \over 5} GM^2/a$ and the
gravitational force on unit mass not outside $a$ is $-{4 \over 3} \pi
G \rho {\bf x}$. The mean square radius of the sphere is $r^2 =
3a^2/5$ so $V = - \left( {3 /5} \right)^{3/2} GM^2/r$. Now
forget real gravity but adopt this form of $V (r)$ for the
potential of one of our extraordinary N-body problems. The force per
unit mass on any particle at ${\bf x}$ is given by
$$-M^{-1} \partial V/\partial {\bf x} = - \left({3\over
5}\right)^{3/2} GM {\bf x}/r^3 = - GM {\bf x}/a^3 = - {4 \over 3} \pi
G \rho {\bf x}$$ where ${\bf x}$ stands for any one of the ${\bf x}_I
- {\overline {\bf x}}$.
Thus for homogeneous spheres this choice of $V (r)$ in our
extraordinary N-body problem exactly mimics the effect of true gravity
both for global radial pulsations of the system and for the forces on
masses within it. However if the system departs from homogeneity this
mimicry is no longer exact. For inhomogeneous spherical systems the
true gravitational potential energy $V (r)$ can always be
written $-k GM^2/r$ with a $k$ that depends on the radial profile. By
taking that to be the $V (r)$ in our extraordinary N-body
problem its Virial theorem will perfectly mimic that of the
gravitational problem but apart from the homogeneous case the forces
on the individual particles of which the system is composed will not
be the same in the mimic. Outside gravitational theory the concept of
an effective potential is widely used in physics and chemistry, e.g.,
in the shell model of the nucleus, in quark-quark interactions at low
energy, and in modelling inter-molecular forces. Now we have shown
that motion in these special potentials can be exactly calculated,
they will no doubt be used as approximations in those applications, as
well as many others. Since the N-body wave-function is known exactly,
so is the correlation energy, but that may not be a useful general
guide to correlation because in our systems the net force on each
particle is directed radially toward the centre-of-mass whatever the
configuration of the other particles may be. Furthermore in many real
systems the interaction between any two particles is strongest when
the particles are closest together while it is weakest for the systems
discussed here. In spite of this it is possible to make systems that
are strongly repulsive when all particles try to come close together
and ones that behave like gravitating systems in the sense that the
overall radius obeys the Virial Theorem for a self-gravitating system.
Even without any repulsion the exclusion principle provides support
for systems of Fermions so with a $V \propto -GM^2/r$ appropriate for
gravity, we find configurations of White Dwarf type.
The N-body problems discussed here arose by direct generalisation of
Newton's work and so skipped the developments of the intervening
centuries. We may nevertheless see how they fit into those
developments. Liouville (1855) showed that if a system of $D$ degrees
of freedom had $D$ integrals of the motion whose mutual Poisson
Brackets vanished, then the remaining $D$ integrals of the motion
could be found as quadratures. He also discovered a large class of
such separable systems while St\"ackel (1890) proved his necessary and
sufficient conditions for separability. Whittaker (1904) gives a good
description of those works while he, Eddington (1915) and Eisenhart
(1934) helped to determine and classify such systems; De Zeeuw (1985)
gives a good historical introduction in his thesis paper. Lynden-Bell
(1962) and Hall (1985) developed different ideas for finding classes
of systems with integrals or configuration invariants. Carter (1968)
extended such results to the motion of charged particles in magnetic
fields in General Relativity. Marshall and Wojciechowski (1988)
determined those potentials in $D$ dimensions for which the motion of
a classical particle separates in suitable coordinates and the
hyper-spherical potential of the systems discussed here can be viewed
as a highly degenerate member of their general $D$-dimensional
ellipsoidally separable potentials. Evans (1990, 1991) has explored
systems that are superintegrable, having more than $D$ integrals for
$D$ degrees of freedom. They separate in several different coordinate
systems and the integrals are the separation constants. This is the
case for our hyperspherical systems (see Appendix).
In all these works separability was achieved by changing the
coordinates only. The idea that separability might be achievable only
via canonical transformations involving the momenta as well as the
coordinates was not exploited. Thus Kovalevski's top (1888) provided
an unexpected new system in which the separation was not of the
standard type. Linear soluble systems were known which were not of the
simple separable type and Routh's (1877) Adam's prize essay on the
Stability of Motion gives a very thorough discussion. Simple examples
of both linear (Freeman 1966) and non-linear problems (Vandervoort
1979, Contopoulos \& Vandervoort 1992) that need momentum dependent
transformations arose in stellar dynamics but it was only recently,
e.g., in the work of Sklyanin (1995) that more general ways of looking
for such systems were found.
Meanwhile a whole body of work based around Lax (1968) operator pairs and the
Inverse Scattering method showed that there were many previously
unsuspected exact solutions in both classical and quantum mechanics.
This field of endeavour is too large to be reviewed here so the reader
is referred to the review volumes Solitons (Bullough \& Caudrey
eds., 1980), Dynamical Systems VI Integrable Systems (Arnold ed.,
1995) and Soliton Theory: a survey of results (Fordy ed., 1990).
The connections between soluble models of N-body problems and field
theory are discussed in the book edited by Bazhanov \& Burden (1995).
In that volume quantum and classical integrable lattice models in one
dimension are considered by Bullough \& Timonen (1995) while two
dimensional models in statistical mechanics are discussed by Baxter
(1995).
Prominent among many exactly soluble N-body models in one dimension
are the Toda lattice (Toda 1967, 1980) also discussed by Henon (1974)
and the Calogero model and its generalisations. [Calogero (1971),
Sutherland (1971), Bullough \& Caudrey (1980), Olshanetsky \&
Perelomov (1995).] In two dimensions Baxter's book (1982) and article
(1995) contain much of interest and certain solvable models in 3
dimensions have been proposed by Baxter (1986) and Bazhanov \& Baxter
(1992, 1993). Probably the most prominent soluble field theory in 2
dimensions is that of Davey-Stewartson (1974), see Anker \& Freeman
(1978), the quantum version of which is considered in Pang et al.
(1990).
Sklyanin (1995) holds out the hope that all these soluble systems may
eventually be seen as special cases of the method of Separation of
Variables and produces some supporting evidence.
Although the above models of interacting systems of many Fermions or
Bosons can be solved exactly in one or two dimensions, the calculation
that follows may be the only exactly solved non-trivial three
dimensional N-body system yet known. Furthermore suitable choices of
the function $V(r)$ will allow a study of the way the form of
interaction (albeit one of our strange global type) affects
Bose-Einstein condensation. It should also prove possible by these
methods to study the effect of rotation on the condensation. However
this paper is solely concerned with the solutions of Schr\"odinger's
equation with the correct symmetry in the wave function, so the
statistical mechanics and Bose-Einstein condensation displayed by
these models is not further discussed here. With the somewhat more
realistic $\delta$ function interaction it has been studied previously
for one dimensional chains and their continuum limits, see, e.g.,
Bogoliubov et al. (1994) and Bullough \& Timonen (1998). It is of
course the case that all soluble models are exceptional and a good
example of the intricacies of non-soluble models was furnished by
Henon (1969).
\section{N-body Solutions of Schr\"odinger's Equation}
Let $m_I$ be the mass of the $I^{\rm th}$ particle and ${\bf x}_I$
its position vector. Writing $M = \Sigma m_I$ for the total mass, the
centre of mass is given by
$${\overline {\bf x}} = \Sigma \mu_I {\bf x}_I \ , \eqno (2.1)$$
where $\mu_I = m_I/M$, which implies
$$\Sigma \mu_I = 1 \ . \eqno (2.2)$$
We define `mass weighted' coordinates relative to the centre of mass \break
${\bf r}_I = \mu^{1/2}_I \left( {\bf x}_I - {\overline {\bf x}}
\right)$ and an associated 3$N$ dimensional vector, ${\bf r}$, in the
space of all the ${\bf r}_I$ by
$${\bf r} = \left( {\bf r}_1 , {\bf r}_2, {\bf r}_3 \ldots
{\bf r}_N \right) \ . $$
The length of ${\bf r}$ is the mass weighted r.m.s. radius of the
system since
$$r^2 = \Sigma \mu_I \left( {\bf x}_I - {\overline {\bf x}} \right)^2 \
. \eqno (2.3)$$
\vfill
\eject
\noindent
This expression may be rewritten in terms of the mutual separations of
the particles since
$$\displaylines{r^2 = \sum_I \mu_I \left( {\bf x}_I - {\overline {\bf
x}} \right) \cdot {\bf x}_I = \sum_I \sum_J \mu _I \mu _J \left( {\bf
x}_I - {\bf x}_J \right) \cdot {\bf x}_I = \cr \hfill{} = \sum_J
\sum_I \mu_J \mu_I \left ( {\bf x}_J - {\bf x}_I\right) \cdot {\bf x}_J \
,\hfill{}\cr} $$
and by adding the last two expressions and halving
the result
$$r^2 = {\textstyle {1 \over 2}} \sum_I \sum_J \mu_I \mu_J ({\bf x}_I - {\bf
x}_J)^2 = \sum_{\ \ I \ <} \! \sum_{\! \! \! J} \mu_I \mu_J ({\bf x}_I
- {\bf x}_J)^2 \ . \eqno (2.4)$$
In practice the ${\bf r}$ vector is constrained by the fact that the
centre of mass is at the origin so
$$\Sigma \mu^{1/2}_I {\bf r}_I = \Sigma \mu_I ({\bf x}_I - {\overline
{\bf x}}) = 0 \ . \eqno (2.5)$$
We define three mutually orthogonal unit vectors ${\hat {\bf X}}$,
${\hat {\bf Y}}$ and ${\hat {\bf Z}}$ in our 3$N$ space by
$$
{\hat{\bf X}} = \left(\mu^{1/2}_1,\, 0,\, 0,\, \mu^{1/2}_2 ,\, 0 ,\, 0
\ldots \mu^{1/2}_N, \, 0 , \, 0\right)$$
$$
{\hat{\bf Y}} = \left(0,\, \mu^{1/2}_1,\, 0,\, 0,\, \mu^{1/2}_2,\, 0 \ldots 0,\,
\mu^{1/2}_N,\, 0\right) \eqno (2.6)$$
$$
\ \ \ {\hat {\bf Z}} = \left(0,\, 0,\, \mu^{1/2}_1,\, 0,\, 0,\, \mu^{1/2}_2 \ldots 0,\,
0,\, \mu^{1/2}_N \right)\ . $$
Then the constraints (2.5) can be rewritten
$${\hat {\bf X}} \cdot {\bf r} = {\hat {\bf Y}} \cdot {\bf r} = {\hat
{\bf Z}} \cdot {\bf r} = 0 \eqno (2.7)$$
which show that ${\bf r}$ is confined to three hyperplanes through the
origin. Defining ${\hat {\bf r}} = {\bf r}/r$ then ${\hat {\bf r}}$
lies on the unit 3$N$ sphere $|{\hat {\bf r}}|^2 = 1$ but the ${\hat
{\bf X}}$ constraint confines it to the intersection of that sphere
with the hyperplane ${\hat {\bf X}}\cdot {\bf r} = 0$, which is a
sphere in 3$N$-1 space; similarly the ${\hat {\bf Y}}$ constraint
leaves it on the intersection of that 3$N$-1 sphere with the
hyperplane ${\hat{\bf Y}} \cdot {\bf r} = 0$, which is a 3$N$-2 sphere
and the third constraint leaves it on the 3$N$-3 sphere orthogonal to
${\hat {\bf X}}$, ${\hat{\bf Y}}$ and ${\hat {\bf Z}}$.
We are concerned with the N-body problems whose potential energies,
$V$, are functions of the magnitude $r$ only, so Schr\"odinger's
equation takes the form
$$-{\textstyle{1 \over 2}} \hbar^2 \sum_I m^{-1}_I {\partial^2 \psi
\over \partial {\bf x}_I \cdot \partial {\bf x}_I} + V \psi = E_T \psi
\ . \eqno (2.8)$$
The key to solving this problem lies in the right choice of
coordinates. In what follows upper case indices run over particle
labels while lower case indices run over coordinate-vector components.
\subsection{Separation of ${\overline x}$}
Let $R_{ij}$ be an orthogonal unit $3N \times 3N$ rotation matrix
which rotates the basis vectors of our $3N$ space so that ${\hat{\bf
X}}$, ${\hat{\bf Y}}$ and ${\hat{\bf Z}}$ are the last three of the
new orthogonal basis vectors. Thus with two alternative notations and
assuming the summation convention over lower case indices only,
$$q_i = R_{ij} r_j = \sum_J {\bf R} _{iJ} \cdot {\bf r}_J \ , \eqno
(2.9)$$
with
$$q_{3N-2} = X \ , \ q_{3N-1} = Y \ , \ q_{3N} = Z \ . \eqno (2.10)$$
Note
$$R_{ij}R_{kj} = \delta _{ik} \ {\rm and} \ \sum_J {\bf R}_{iJ} \cdot
{\bf R}_{kJ} = \delta_{ik} \ . \eqno (2.11)$$
Let $a$ run from 1 to 3$N$-3 (rather than from 1 to $3N$). Then the
$q_a$ together with the coordinates ${\overline {\bf x}}$ form a
complete set of independent orthogonal coordinates for our system.
We shall need the partial derivatives from (2.1) and below (2.2)
$$\partial {\overline{\bf x}}/\partial {\bf x}_I = \mu_I
{\underline{\mbox{\boldmath$ \delta$}}} \ , \eqno (2.12)$$
$$\partial {\bf r}_J/\partial {\bf x}_I = \sqrt{\mu_J} \left
( \delta_{JI} - \mu_I \right) {\underline{\mbox{\boldmath$ \delta$}}} \ ,
\eqno (2.13)$$
where $
{\underline{\mbox{\boldmath$ \delta$}}}$ is the unit $3 \times 3$
matrix. The centre of mass motion will separate so our wave functions
may be taken in the form $\psi = {\overline \psi} ({\overline {\bf
x}}) \widetilde \psi (q _a \ldots )$ so
$${\partial \psi \over \partial {\bf x}_I} = {\partial {\overline
\psi} \over \partial {\overline {\bf x}}}\, \mu_I \widetilde \psi +
{\overline \psi} \, {\partial \widetilde \psi \over \partial {\bf x}_I} \ .
\eqno (2.14)$$
For Schr\"odinger's equation we shall need
$$\sum_I \mu^{-1}_I\, {\partial ^2 \psi \over \partial {\bf x}_I \cdot
\partial {\bf x}_I} = {\partial^2 \overline \psi \over \partial
{\overline{\bf x}} \cdot \partial {\overline {\bf x}}} \, \widetilde
\psi + \overline \psi \Sigma \mu^{-1}_I {\partial^2 \widetilde \psi
\over \partial {\bf x}_I \cdot {\bf x}_I} \ , \eqno (2.15)$$ where the
cross derivative term has vanished because $\widetilde \psi$ only
involves differences of the coordinates $x_I$ so $\Sigma
\partial \widetilde \psi/\partial {\bf x}_I = 0$, i.e., $\widetilde
\psi$ is independent of where the system-as-a-whole is.
To evaluate the second term we need from (2.9) and (2.13)
$${\partial \widetilde \psi\over \partial {\bf x}_I} = \sum_K {\partial
\widetilde \psi \over \partial q_j} \, {\partial q_j \over \partial
{\bf r}_K} \cdot {\partial {\bf r}_K \over \partial {\bf x}_I} =
{\partial \widetilde \psi \over \partial q_j} \left( {\bf R}_{jI}
\sqrt{\mu_I} - \sum_K \sqrt{\mu_K} \mu_I {\bf R}_{jK} \right) \
. \eqno (2.16)$$
We check that indeed $\Sigma \partial \widetilde \psi / \partial {\bf x}_I =
0$ by summing this over $I$ and noting that the two sums cancel because $\Sigma \mu_I = 1$. We now proceed to the last term in
Schr\"odinger's equation (2.15)
$$\sum_I {1 \over \mu_I}\, {\partial^2 \widetilde \psi \over \partial
{\bf x}_I \cdot \partial {\bf x}_I} = {\partial^2 \widetilde \psi
\over \partial q_\ell \partial q_j }\, \sum_I {\bf R}_{jI} \cdot {\bf
R}_{\ell I} - \sum_K \sum_I {\partial^2 \widetilde \psi \over \partial q_j
\cdot \partial {\bf x}_I} \sqrt{\mu_K} {\bf R}_{jK} \ .$$
But the last term involves $${\partial \over \partial q_j} \left
( \sum_I {\partial \widetilde \psi \over \partial {\bf x}_I} \right)$$
which is zero and $$\sum_I {\bf R}_{jI} \cdot {\bf R}_{\ell I} =
\delta_{j\ell}$$ because $\bf R$ is an orthogonal matrix. Hence we have the
desired expression
$$\sum_I \mu^{-1}_I {\partial^2 \psi \over \partial {\bf x}_I \cdot
\partial {\bf x}_I} = {\partial^2 {\overline \psi} \over \partial
{\overline {\bf x}} \cdot \partial {\overline {\bf x}}} {\widetilde
\psi} + {\overline \psi} {\partial^2 {\widetilde \psi} \over \partial
q_a \partial q_a} \ . \eqno (2.17)$$
$a$ has replaced $j$ in the final term because $\widetilde \psi$
is only dependent on the first 3$N$-3 of the $q_j$, and we remember
that $q_aq_a = r^2$ since $X$, $Y$ and $Z$ are all zero.
On division by $\psi$ Schr\"odinger's equation now takes the form
$$-{\textstyle {1 \over 2}} \hbar^2 M^{-1} \left( {1 \over {\overline
\psi}} \, {\partial^2 {\overline \psi} \over \partial {\overline {\bf
x}} \cdot \partial {\overline {\bf x}}} + {1 \over {\widetilde \psi}}
\, {\partial^2 {\widetilde \psi} \over \partial {\bf q} \cdot \partial
{\bf q}} \right) + V(r) = E_T \ , \eqno (2.18)$$
where ${\bf q}$ stands for the $3(N-1)$ vector $q_a$. The equation
clearly separates with the final three terms dependent on the $q_a$
only and the first dependent on ${\overline {\bf x}}$ only, so it must
be constant. Without loss of generality we can take the total
momentum to be $\hbar {\bf K}$. Then ${\overline \psi} = \exp (i{\bf
K} \cdot {\overline {\bf x}})$ and writing $E = E_T - {1 \over 2}
\hbar^2 K^2/M$ we find
$$- {\hbar^2 \over 2M} \, {\partial^2 {\widetilde \psi} \over \partial
{\bf q} \cdot \partial {\bf q}} + V(r) {\widetilde \psi} = E
{\widetilde \psi} \eqno (2.19)$$
where ${\bf q} \cdot {\bf q} = r^2$.
\subsection{Separation of Angular Coordinates}
Equation (2.19) clearly separates again in hyperspherical polar
coordinates but it is simplest to write them symbolically by putting
${\bf q} = r {\hat {\bf r}}$ and regarding $r$ as independent of the
angular coordinates ${\hat {\bf r}}$. We need the partial
differentials
$$\partial r/\partial {\bf q} = {\hat {\bf r}} \eqno (2.20)$$
$$\partial {\hat {\bf r}}/ \partial {\bf q} = \partial /\partial {\bf
q} \left( {\bf q}/r\right) =
r^{-1} ({\underline {\mbox {\boldmath $\delta$}}} - {\hat {\bf r}}
{\hat {\bf r}} ) \eqno (2.21)$$
and by writing ${\hat {\bf r}} = {\bf q}/r$ and using (2.20)
$$ \partial / \partial {\bf q} \cdot {\hat {\bf r}} = r^{-1} (3N-4) \
. \eqno (2.22)$$
We write ${\widetilde \psi} = \psi_r (r) {\hat {\psi}} ({\hat {\bf r}})$ and notice
that ${\hat \psi}$ is constant on radial lines so that ${\hat {\bf r}}
\cdot \partial {\hat \psi}/\partial {\hat {\bf r}} = 0$. Then
$${\partial {\widetilde \psi} \over \partial {\bf q}} = {\partial \psi_r \over
\partial r} \, {\hat {\bf r}} {\hat \psi} + r^{-1} \psi_r {\partial {\hat \psi}
\over \partial {\hat {\bf r}}} \eqno (2.23)$$
and
$${\partial^2 {\widetilde \psi} \over \partial {\bf q} \cdot \partial {\bf q}} = {\partial
^2 \psi_r \over \partial r^2} \, {\hat \psi} + (3N-4) r^{-1} {\partial \psi_r
\over \partial r} \, {\hat \psi} + r^{-2} \psi_r {\partial^2 {\hat \psi} \over
\partial {\hat {\bf r}} \cdot \partial {\hat {\bf r}}} \eqno (2.24)$$
dividing by $\psi$ Schr\"odinger's equation now takes the form
$$\psi^{-1}_r \left[ r^2 \partial ^2 \psi_r / \partial r^2 + (3N-4)r
\partial \psi_r/\partial r \right] - \alpha^2 r^2 + U(r) r^2 = {\hat
{\psi}}^{-1} A ({\hat {\psi}}) \ . \eqno (2.25)$$ $\alpha$ is given by
$\alpha^2 = -2ME \hbar^{-2}$ and $U(r) = -2MV(r) \hbar^{-2}$. The
angular operator $A$ (the hyper-angular-momentum operator) is given by
$$A ({\hat \psi}) = - \partial^2 {\hat \psi}/\partial
{\hat {\bf r}} \cdot \partial {\hat {\bf r}} \ . \eqno (2.26)$$
The angular operator $A$ also appears in the generalised $\nabla ^2$
in 3$N$-3 dimensions viz
${\partial^2 / \partial {\bf q} \cdot \partial {\bf q}}$,
so we shall first study the hyper-spherically symmetric solutions of
$\partial^2 \chi / \partial {\bf q} \cdot \partial {\bf q} = 0$.
Evidently
$${d^2 \chi \over dr^2} + {(3N-4) \over r} \, {d \chi \over d r} = 0$$
so $d \chi/dr = Cr^{-(3N-4)}$ therefore $\chi = B r^{-(3N-5)}$ where
we have omitted a constant of integration which is irrelevant to our
purpose. We see this is correct by considering the first non-trivial
case in which terms other than ${\overline {\bf x}}$ are involved
which is $N=2$. Then the space of the $q_a$ is 3 dimensional and the
elementary solution has $\chi = Br^{-1}$. Now our generalised
$\nabla^2$ knows no particular origin so if $\chi$ is a solution for
$r \neq0$ then so is ${\widetilde \chi} = B |{\bf r} - {\bf r}_0 |
^{-(3N-5)}$ for ${\bf r} \neq {\bf r}_0$. We expand such solutions
both for $|{\bf r}| < |{\bf r}_0|$ and for $|{\bf r}| > |{\bf r}_0|$ in powers
and the coefficients of these powers are $Y_L ({\hat {\bf r}})$
hyperspherical harmonics, just as they are in 3 dimensions. In
particular the $L^{\rm th}$ harmonics have a power in $r$ of either
$r^L$ or $r^{-L-(3N-5)}$. Looking for solutions of the form
$r^Lf({\hat {\bf r}})$ to our generalised $\nabla^2 =0$ we see that in $D=3N-3$ dimensions
$$r^2 {d^2 \over dr^2} \left(r^Lf\right) + (D-1) r {d \over dr}
\left( r^Lf \right) = r^L A(f) \ ,$$
i.e.,
$$L(L + D-2)\, f = A(f) \eqno (2.27)$$
and hence the eigenvalues of $A(f)$ are $L(L+3N-5)$.
Notice that for a two body problem this reduces to the $L (L+1)$ in 3
dimensions that we know so well. We shall return later to look for
the degeneracies of these different eigenstates but for solving
Schr\"odinger's equation the eigenvalues are sufficient. For the
detailed separation of the $3N-4$ angular coordinates each in turn see
the Appendix.
\subsection{The Radial Equation}
Schr\"odinger's equation (2.25) now reads
$$d^2 \psi_r/dr^2 + (3N-4)r^{-1} d\psi_r/dr - [\alpha^2 - U(r) +L
(L+3N-5)r^{-2}] \psi_r = 0 \ . \eqno (2.28)$$
Now in the corresponding classical N-body problem we showed that the
solution for the radial pulsations of the whole N-body system could
be found in terms of the radial pulsation of the corresponding two
body problem with the same $U(r)$. With $N=2$ we have the usual
Schr\"odinger equation for a spherical potential with angular momentum
$\ell$
$$d^2 \psi_2/dr^2 + 2r^{-1}d\psi_2/dr - (\alpha^2 - U(r) + \ell (\ell
+1) r^{-2}) \psi_2 = 0 \ . $$
We shall suppose that this problem has been solved and the
corresponding eigenvalues and eigenfunctions are known. Now put
$\psi_2 = r^{\beta} \chi_2$; then $\chi_2$ obeys
$$\chi_2 '' + (2 \beta +2) r^{-1} \chi'_2 - \left[ \alpha^2 - U(r) + \ell
(\ell + 1) - \beta (\beta +1) \right] \chi_2 = 0 \ . \eqno (2.29)$$
Writing $\ell (\ell +1) - \beta (\beta +1) = (\ell - \beta) (\ell + \beta
+1)$ one notices that putting
$$\beta = {\textstyle {3 \over 2}} (N-2) \eqno (2.30)$$
and $\ell = L+
\beta$,
transforms equation (2.29) into precisely (2.28). We deduce the
surprising theorem below:
{\em Theorem}: If the energy levels of the usual Schr\"odinger
equation in 3 dimensions are $E(P, \ell)$, where $P= 1,2,3 \ldots$ is
the radial quantum number and $\ell$ is the angular momentum quantum
number, then the energy levels of the N-body problem with the `same'
`potential' $U(r)$ are $E \left(P, L+ {\textstyle {3 \over 2}}
(N-2)\right)$. Furthermore the radial parts of their wave functions
are related by $\psi_r (P, L, r) = r^{-{3 \over 2} (N-2)} \psi_2 \left(P, L
+ {3 \over 2} (N-2), r \right)$.
Notice that the ground states are not the same because $L+{3 \over
2}(N-2)$ can not be zero for $N>2$. Thus the ground state of the
N-body system corresponds to an $\ell$ of ${3 \over 2} (N-2)$ which
can be a high angular momentum state of the two body problem.
We see at once that the N-body problem will have proper energy levels
even if strongly {\em attractive} $r^{-2}$ potentials are added to $U$
just because of this effective increase in the $\ell$ of the ground
state.
The theorem above is very powerful in that it enables us to use
everything that is known about the solutions to the normal
Schr\"odinger equation in spherical potentials and transform it into
knowledge of our N-body problems. As is well known, not only the
generalised Kepler potential $U(r)\! =\! \zeta r^{-1}\! -\! \zeta_2
r^{-2}$, but also similarly generalised square well potentials, and
the $U \propto \delta (r-r_0)$, $\delta$-function potentials as well
as oscillator potentials $U(r) = -{1 \over 2} \kappa^2 r^2 - \zeta_2
r^{-2}$ all have pretty solutions for the eigenvalues and
eigenfunctions. All this takes over directly. However, when
$\zeta_2$ is so negative that the two body problem has no ground state
there is perhaps room for doubt as to whether we can use that
potential's higher angular momentum states for our N-body problem.
To allay any such doubts and provide a pretty illustration of the
truth of our general theorem, we now solve the N-body equation for the
generalised Kepler potential.
\subsection{Generalised Kepler Problem}
We have already shown that the wave function takes the form
$$\psi = \exp (i {\bf k} \cdot {\overline {\bf x}}) Y_L ({\hat {\bf
r}}) \psi_r (r)$$
where $Y_L$ may be given a further 3$N$-5 indices corresponding to
various components of $\bf L$. We must now determine $\psi_r$.
In solving equation (2.28) we follow the standard method of solution
beautifully laid out in the book by Pauling \& Wilson (1935). Setting
$\alpha r = \tilde r$ and $\zeta/\alpha = \tilde \zeta$ and keeping
only the terms that dominate at large $\tilde r$ we find
$$d^2 \psi_r / d \tilde r^2 - \psi_r \sim 0\ .$$
So the asymptotic solutions behave as $\exp \pm \, \tilde r$. Only the
minus sign is acceptable so we write $\psi_r = \eta (\tilde r) \exp (-
\tilde r)$ and derive the equation for $\eta$ valid for all $\tilde
r$,
$$\eta '' + (3N\! - 4) \tilde r^{-1} \eta ' - 2 \eta' + \biggl \{ \left
[ {\tilde \zeta} - (3N\! \! - \! 4) \right] {\tilde r}^{-1}\! - \!
\left[ \zeta _2 +L (L+3N\! \! -\! 5) \right] {\tilde r}^{-2} \biggr\}
\eta = 0 \; . \eqno (2.31)$$
We look for power series solutions of the form
$$\eta = \sum^\infty_{p=0} a_p \tilde r ^{s+p}$$with $a_0 \neq 0$ and
find the recurrence relation
$$\displaylines{ \biggl\{ (p+s) (p + s + 3N -5) - \bigl[ \zeta_2 +L
(L+3N-5) \bigr] \biggr\} a_p = \cr \hfill = \left[ 2(p+s) + 3(N-2) -
\tilde \zeta \right] a_{p-1} \ . \hfill {(2.32)}}$$ The indicial
equation has $a_{-1} = p = 0 $ and yields a quadratic equation for $s$
$$s^2 + (3N -5) s - \left[ \zeta_2 + L (L+3N-5) \right] = 0 \ . \eqno
(2.33)$$
In the pure Kepler case with $\zeta_2=0$ this yields $s = L$ or
$-(L+3N-5)$ of which only the positive $s=L$ solution obeys the
boundary condition at the origin. For general $\zeta_2$ the solutions
are (cf (2.30))
$$s = -{\textstyle {1 \over 2}} (3N-5) \pm \sqrt{ \left[ L+{\textstyle
{1 \over 2}} (3N-5) \right]^2 + \zeta_2} = - {\textstyle {1 \over 2}}
- \beta \pm \sqrt{ \left( {\textstyle{1 \over 2}} + L + \beta
\right)^2 + \zeta_2} \ , \eqno (2.34)$$
of which only that with the $+$ sign obeys the boundary condition at
$r = 0$. When $\zeta_2 < 0$ a more detailed discussion is given
later. If the series for $\eta$ does not terminate the asymptotic
form of the recurrence relation gives $a_p \simeq 2 a_{p-1}/p$ so
$\eta \propto e^{2 \tilde r}$ and $\psi_r$ is divergent at $\infty$;
so the series must terminate at $a_{P-1}$ say and in (2.32)
$$\tilde \zeta = 2 (P+s) + 3(N-2)=2(P+s+\beta ) \ , \eqno (2.35)$$
with $s$ given by taking the upper sign in (2.34) (i.e., $s = L$ when
$\zeta_2 = 0$).
Remembering that $\tilde \zeta = \zeta/\alpha$ and that $\alpha^2 = -2
ME \hbar ^{-2}$ expression (2.35) can be recast as the energy spectrum
$$E = - {\hbar ^2 \over 8M} \, {\zeta^2 \over (P + s + \beta)^2} = -
{\hbar^2 \over 2M} \, {\zeta^2 \over \left[2P-1 + \sqrt{ \left(2L
+3N-5 \right)^2 + 4 \zeta_2} \right]^2} \ . \eqno (2.36)$$ In
accordance with our theorem the energy levels with general $N$ are
given by putting $N=2$ and then replacing $L$ by $L+\beta = L + {3
\over 2} (N-2)$. Of course if $\zeta_2=0$ we have $s=L$ and the
theorem is then obvious from the first form. Notice that the theorem
really applies to $\alpha^2 = -2ME/\hbar^2$ thus we can only apply it
to $E$ itself if we consider a two body problem with the same mass $M$
as the N-body problem; $\zeta^2$ is also taken as unchanged since we
require both problems to have the same $U(r)$. However this in no
way restricts us to N-body problems with $M$ and $\zeta$ independent
of $N$; it merely means that we change correspondingly the masses $M$
and coefficients $\zeta$ in the two body problems with which we
compare N-body problems of different $N$.
Some may wish to see the precise Schr\"odinger hydrogen atom spectrum
with the correct reduced mass emerging when $N=2$; to get this we must
evaluate $\zeta$ in terms of $Ze^2$. Our potential energy is $V =
-{\textstyle {1 \over 2}} \hbar^2 M^{-1} \zeta/r$ but this $r$ is not
the separation of the nucleus from the electron, $R$, but the mass
weighted r.m.s. separation of them from the centre of mass. Hence $r
= (mm_n/M^2)^{1/2} R$ where $m$ and $m_n$ are the masses of the
electron and the nucleus respectively. Setting $V = - {Ze^2 \over R}$
we deduce that $\zeta = 2 (mm_n)^{1/2} \hbar ^{-2} Ze^2$. Inserting
this $\zeta$ into (2.36) along with $N=2, \zeta_2 = 0, n = P+L, M =
m+m_n$ and putting the reduced mass $mm_n/M = m_r$ the energy levels
of hydrogenic atoms are given by
$$E = - {m_r \over 2 \hbar^2} \, {(Ze^2)^2 \over n^2} $$
just as they should be.
We now return to the question of how negative $\zeta_2$ can be. Since
$L$ can be zero the energy of the ground state ceases to
be real if $\zeta_2 < - \left( {3N-5 \over 2} \right)^2$ which gives $\zeta_2 < - {1 \over 4}$ for the familiar
case $N=2$. Such strongly attractive forces cause the particles to
propagate into the nucleus and the ground state ceases to exist. It
may be seen that the limiting case has a wave function $\psi \propto
r^{- {1 \over 2}}$ near the origin which is easily still square
integrable $\int \psi \psi ^\star r^{2} dr < \infty$. This is also
true for the limiting case $\zeta_2 = - \left( {3N-5 \over 2} \right) ^2$ for then $\psi
\propto r^{-{1 \over 2}(3N-5)}$ and $\int \psi \psi ^\star r^{3N-4} dr
< \infty$. The limits are surpassed for the attractions of magnetic
monopoles on the magnetic moments of protons and for charged monopoles
attracting spinning electrons (Lynden-Bell \& Nouri-Zonoz 1998).
\section{Level Degeneracies}
In equations (2.26) and (2.27) and the attendant discussion we showed
that the solutions of our Schr\"odinger equation consisted of a
hyperspherical harmonic in $3N-3$ dimensions times a radial function.
Furthermore the hyperspherical harmonics of degree $L$ in $D$
dimensions are the coefficients of $r^L$ in the polynomial solutions
of Laplace's equation in $D$ dimensions. Thus the degeneracy of the
states of given $L$ and given radial quantum number $P$ will be equal
to the number of independent polynomial solutions of Laplace's
equation of degree $L$ in $D=3N-3$ dimensions (i.e., harmonic
polynomials). To determine this number we first ask how many
independent polynomials of degree $L$ exist in $D$ dimensions without
the harmonic requirement. Each can be considered as a term of the
form $\Pi _i x_i^{l_i}$ where $i$ runs from 1 to $D$ and $\Sigma l_i
=L$. That number of polynomials is equal to the number of ways of
dividing $L$ objects into $D$ groups where a group is allowed to
contain no objects. If we take $L$ units and $D-1$ dividing bars then
the number of ways of ordering them is $(L+D -1)!$ and if we disregard
the ordering of the $D-1$ bars among themselves and the $L$ units
among themselves the number of ways of sorting them into groups is
$$G (L,D) = {(L+D-1)! \over (D-1)!L!}\ , \eqno (3.1)$$ so this is the
number of independent polynomials of degree $L$. Let $f_L ({\bf
x}_a)$ be such a homogeneous polynomial of degree $L$ in the ${\bf
x}_a$. In general
$$\nabla^2 f_L = \sum_a {\partial \over \partial {\bf x}_a} \cdot {\partial \over \partial {\bf x}_a} \, f_L = f_{L-2}$$ where $f_{L-2}$ is such a polynomial of degree $L-2$ which will have $G(L-2, D)$ independent coefficients. The condition that $f_L$ be harmonic $(\nabla^2f_L = 0)$ thus imposes $G(L-2, D)$ constraints on the $G(L, D)$ free coefficients in $f_L$.
Thus the number of independent harmonic polynomials of degree $L$
in $D$ dimensions is
$$\displaylines{\ g (L,D) = G(L,D) - G(L-2,D) = \hfill \cr \hfill{}\cr
\hfill \quad \ \, = {(L+D-3)! \left [ (L+D-1)(L+D-2)-L(L-1)\right] \over
(D-1)!L!} = \hfill (3.2) \cr \hfill{}\cr = {(L+D-3)! \over (D-2)!L!}
(2L+D-2) \ . \qquad \qquad \qquad \qquad \qquad \quad \ \ \, }$$
Notice that for the familiar case $D=3$ this gives the correct answer,
$2L+1$, for the degeneracy of the states of given $L$.
When $\zeta_2=0$ we have the extra degeneracy between the $s, p, d, f$
levels of hydrogen. Then a state of principal quantum number $n$ can
be obtained by combining a wave function of given $L$ with a radial
wave function with radial quantum number $P = n - L \geq 1$. Thus, to
find the degeneracy of states with a given $n$, we need to know the
number of harmonic polynomials of degree less than or equal to $n-1$,
because the $n-L-1$ extra quanta are taken up by different radial wave
functions. To find this number we merely add a dummy group `$o$' to
our sorting of $L$ objects into groups and ignore the number of units,
$n_o$, that falls into that group. Thus the required answer is
$$g_H (n,D) = g (n-1, D+1) = {(n+D-3)! \over (D-1)! (n-1)!} (2n + D-3) \ .
\eqno (3.3)$$
For $D=3$ this reduces to $n^2$, which is the
familiar degeneracy of the $n^{\rm th}$ hydrogen level before spin is
considered. Thus for our problems the degeneracy of levels of
hyper-angular-momentum $L$ for a system of $N$ particles is $g(L,
3N-3)$ with $g\ {\rm given}$ by (3.2), while for $\zeta_2=0$ the degeneracy of
the $n^{\rm th}$ level is \break $g_H (n, 3N-3) = g (n-1, 3N-2)$.
The above degeneracies are worked out assuming that none of the
particles are identical. In practice we are more interested in
problems with $N$ identical Bosons or $N$ identical Fermions and they
require wave functions with even or odd permutational symmetries so we
now study that question.
\section{Symmetry under permutation of particles}
\subsection{Bosons}
For Bosons we need wave-functions that are symmetrical for the
interchange of any two particle labels ${\bf x}_I, {\bf x}_J$. Both
${\overline {\bf x}}$ and $r$ have the required symmetry when the
particles are of equal mass. Just as the magnitude of the angular
momentum treats $x, y$ and $z$ symmetrically in 3 dimensions, so the
magnitude of the hyperangular momentum $L$ is symmetrical for particle
interchange. However the components of the term $L_{ij}$ and the
vector ${\hat {\bf r}}$ are not symmetrical under the interchange of
particle labels. In \sect2 we found the solutions for our
$N$-particle wave functions in the form
$$\psi = \exp (i {\bf k} \cdot {\overline {\bf x}}) Y_L ({\hat {\bf
r}}) \psi _r (r) \ , \eqno (4.1)$$
\vfill
\eject
\noindent
where $\psi_r$ depends only on the scalar magnitude $L$. The only
term that is not automatically symmetrical for particle interchange is
$Y_L ({\hat {\bf r}})$ but even that will be automatically symmetrical
when $L=0$, because $Y_0$ is constant. Thus the ground state and all
$s$-states are automatically symmetrical and are possible states for a
system of identical Bosons.
The states we have been discussing are {\em not} the one-particle
states commonly considered as components of $N$-particle product
states (or, for Fermions, Slater determinants); rather our states are
themselves $N$-particle states. To get a symmetrical $N$-particle
state from one lacking that symmetry we merely add all the wave
functions obtained by permuting the labels on the particles. But
whereas each of our wave-functions thereby generates one boson
$N$-particle state, such a state in general comes from a number of
different unsymmetrical wave-functions so we can no longer count the
degeneracies by the arguments of \sect3. However the arguments of
\sect3 connect the number of $Y_L$ functions with the number of
polynomials that are homogeneous and both of degree $L$ and harmonic.
If we can count interchange symmetric polynomials independent of
${\overline {\bf x}}$ which are homogeneous of degree $L$ and
solutions of Laplace's equation, we have the degeneracy of the quantum
states of hyper-angular-momentum $L$.
Let $F_L ({\bf x}) = F_L ({\bf x}_1, {\bf x}_2, \ldots {\bf x}_N)$ be a
homogeneous polynomial of degree $L$ in ${\bf x}$ which is symmetric
under the interchange of any ${\bf x}_I$ with ${\bf x}_J$. Then \break $F_L
(\lambda {\bf x}_1, \lambda {\bf x}_2, \ldots \lambda {\bf x}_N) =
\lambda^L F_L ({\bf x}_1, {\bf x}_2, \ldots {\bf x}_N) = \lambda^L
F_L ({\bf x}_1, {\bf x}_N, \ldots {\bf x}_2), \ {\rm etc.} $ Consider
$F_L ({\bf x} - {\overline {\bf x}} ) = F_L ({\bf x}_1 - {\overline
{\bf x}}, \ldots {\bf x}_N - {\overline {\bf x}})$. It is also a
homogeneous polynomial of degree $L$ in ${\bf x}$ and is also
symmetric, but it has the property that it is invariant under the
transformation ${\bf x}_I \rightarrow {\bf x}_I + {\mbox {\boldmath
$\Delta$}}$ for all $I$ (because then ${\overline {\bf x}} \rightarrow
{\overline {\bf x}} + {\mbox {\boldmath $\Delta$}}$). Thus such
functions do not depend on the position of the centre of mass.
However it can happen that $F_L ({\bf x} - {\overline {\bf x}})$ is
identically zero even when $F_L ({\bf x})$ is not. For this to happen
$F_L ({\bf x})$ must be of the form $F_L ({\bf x}) = {\overline {\bf x}}
\cdot {\bf F}_{L-1} ({\bf x})$ where ${\bf F}_{L-1} ({\bf x})$ is a
vector each of whose components is a polynomial of degree $L-1$ in
${\bf x}$ which is symmetric under interchange of ${\bf x}_I$ and
${\bf x}_J$. Now let ${\overline G}(L)$ be the number of independent
symmetric polynomials which are homogeneous of degree $L$ in 3$N$
dimensions. Then the number of such polynomials giving rise to
non-zero $F_L ({\bf x} - {\overline {\bf x}})$ will be ${\overline G}
(L)$ less the number of free coefficients in the $- {\overline {\bf
x}} \cdot {\bf F}_{L-1}$ term which we might expect to be $3
{\overline G} (L-1)$. However that is not quite right because a
polynomial with a factor ${\overline {x}}\, \overline {y}$ will occur
as a possibility in both the $x$ and $y$ components of ${\bf F}_{L-1}$
and in subtracting $3 {\overline G} (L-1)$ we will have subtracted its
number of free coefficients not once but twice. The same double
counting will have occurred for polynomials with factors ${\overline
y} \, {\overline z}$ or ${\overline z} \, {\overline x}$ so we should
be subtracting not $3 {\overline G} (L-1)$ but rather $3 {\overline G}
(L-1) - 3 {\overline G} (L-2)$. However even that is not quite
correct because there may be polynomials with a factor ${\overline
x}\, {\overline y}\, {\overline z}$. They will have been subtracted
off three times in $3 {\overline G} (L-1)$ but added back in three
times in $3 {\overline G} (L-2)$ so they will still be there and they
should not be since they clearly vanish when ${\bf x} - {\overline
{\bf x}}$ is written for ${\bf x}$. Thus finally the number of
independent symmetric polynomials which are homogeneous of degree $L$
and independent of ${\overline {\bf x}}$ is
$${\overline G}_1 (L) \equiv {\overline G} (L) - 3 {\overline G} (L-1)
+ 3 {\overline G} (L-2) - {\overline G} (L-3) \ . \eqno (4.2)$$
However, we still have to impose Laplace's equation
$$\sum_I \nabla^2_I F_L \equiv \nabla^2 F_L=0 \ . $$ Now in general
$\nabla^2 F_L ({\bf x} - {\overline {\bf x}})$ will be a polynomial of
degree $L-2$ in ${\bf x}$. However, since $F_L ({\bf x} - {\overline
{\bf x}})$ is invariant to the transformation ${\bf x} \rightarrow
{\bf x} + \mbox {\boldmath $\nabla$}$, $\nabla^2 F_L ({\bf x} -
{\overline {\bf x}})$ will also have that property. Thus in general
we may write $$\nabla^2 F_L ({\bf x} - {\overline {\bf x}}) = F_{L-2}
({\bf x} - {\overline {\bf x}})$$ where $F_{L-2}$ is also symmetrical
for particle label interchange since $\nabla^2$ does not destroy that
property. Thus the condition $F_{L-2} ({\bf x} - {\overline {\bf x}})
\equiv 0$ will put ${\overline G}_1 (L-2)$ constraints on the
${\overline G}_1 (L)$ free coefficients of the homogeneous $L^{{\rm
th}}$ degree polynomial $F_L ({\bf x} - {\overline {\bf x}})$. There
will be just ${\overline G}_1 (L) - {\overline G}_1 (L-2)$ free
coefficients left in $F_L ({\bf x} - {\overline {\bf x}})$ after
imposing the harmonic condition so this is the degeneracy of the $Y_L$
that corresponds to the $(2L+1)$ with $L$ even for the 2 boson
problem. Since ${\overline G}_1$ is known in terms of ${\overline G}$,
we have reduced our problem to that of determining the number of
exchange-symmetric homogeneous polynomials of degree $L$ in 3$N$
dimensions. This is the crux of our problem and it took us
considerable thought to solve it. Exchange symmetry
involves exchanging vectors ${\bf x}_I$ with ${\bf x}_J$, so we do not
need symmetry in all 3$N$ dimensions but only between them taken in
triples. We shall begin our
considerations with the simpler case of $N$ bosons on a line with each
having but one coordinate $x_I$. We then wish to know how many
independent exchange-symmetric polynomials there are which are
homogeneous of degree $L$ in $N$ dimensions.
Let $\Phi = S \Pi_I x^{\ell _I}_I$ be a symmetrical polynomial of
degree $L$ with $N$ factors $x^{\ell _I}_I$. $S$ is the symmetrising
operator which is the sum over all permutations of the particle labels
$I$. Different symmetric polynomials are characterised by different
sets of integers $(\ell_1, \ell_2, \ldots \ell_N)$ or partitions of
the integer $L$ into $N$ parts, some of which may be zero. We
construct the generating function
$$B_1 (u,x) = \sum^\infty_{n=1} \sum^\infty_{\ell = 0} p(n, \ell) u^n
x^\ell$$
where $p(N,L)$ is the number of partitions of $L$ into $N$ integers
that may be zero, and for convenience we have defined $p (0, \ell) =
0$ and $p (n, 0 ) = 1$. We now show how the theory of partitions
allows us to construct the function $B_1 (u, x)$.
\subsection{Partitions of an integer $L$}
We learn from Abramowitz \& Stegun (1964) that `The number of
decompositions of an integer $L$ into integer summands without regard
to order is called $p (L)$'.
For example, five may be written
$$5 = 4+1 = 3+2 = 3+1+1 = 2+2+1 = 2+1+1+1 = 1+1+1+1+1 \ , \eqno
(4.3)$$
so we deduce that $p(5) =7$. It is easiest to work with the
generating function for the $p(L)$ which we call $A(x)$. For this
there is a standard result see, e.g., Hardy \& Ramanujan (1918),
$$A(x) = 1 + \sum^\infty_1 p(L) x^L = \prod\limits^\infty_{\ell =1} (1 -
x^\ell)^{-1} \ . \eqno (4.4)$$
\vfill
\eject
For what follows it is essential to understand how this standard
result comes about. To do so we rewrite the product by expansions in
powers of $x$
$$\begin{array}{lccc}
A(x) = & \! \! \left( 1 + x^1 + x^2 + \ldots \right) & \! \! \left( 1 + (x^2)^1 +
(x^2)^2 + \ldots \right) & \! \! \left( 1 + (x^3)^1 + (x^3)^2 \ldots
\right) \cr
&({\rm generates \ units}) & ({\rm generates \ twos}) & ({\rm generates
\ threes })
\end{array}$$
\vskip -0.3cm
\rightline {(4.5)}
\vskip -0.6cm
$$\hfill \begin{array}{ccc}
\left( 1+ (x^4)^1 + \ldots \right) & \left( 1 + (x^5)^1 + \ldots \right)
&(\ldots) \cr
({\rm generates \ fours}) & ({\rm generates \ fives})& \qquad \qquad \qquad .
\end{array}$$
To see how the coefficient of $x^5$ in this expression is $p(5)=7$ we
first notice that we must take the 1 from all brackets after the fifth,
since otherwise we would get too high a power of $x$. In the 5$^{\rm
th}$ bracket we can take the $x^5$ term but then we must take the 1
from all earlier brackets. Alternatively we take the 1 in the 5$^{\rm
th}$ bracket. In the latter case we turn to the fourth bracket. Here
we may take the $x^4$ term but that can only be coupled to the $x$
term in the first bracket in which case we get the split of 5 into
$4+1$. Turning now to the third bracket and taking the $x^3$ term we
can take it with either the $(x^2)^1$ bracket of the second term to
yield $3+2$ or with the $(x^1)^2$ term of the first bracket to yield
$3+1+1$. Similarly from the second bracket we could take the
$(x^2)^2$ term with the $x$ from the first bracket to give $2+2+1$ or
the $(x^2)^1$ term with the $x^3$ from the first to give $2+1+1+1$.
Finally we could take the $x^5$ from the first bracket to give
$1+1+1+1+1$. Thus the first bracket yields the number of ones in the
sum, the second the number of twos, the third the number of threes,
etc., and in this way the coefficient of $x^L$ yields $p(L)$.
However, we need the restricted partition of $L$ into $N$ or fewer
non-zero integers $p(N,L)$. These are sums of partitions $p_1(N,L)$
into exactly $N$ non-zero integers. Looking at (4.3) we see, for
example, that $p_1(2,5) = 2 = p_1 (3,5)$. If we place a factor $u$
along with each factor $x^{\ell}$ in (4.5), then the power of $u$ in
each term will tell us how many parts there are in the partition
generated by a particular term. Thus in place of $A(x)$ we consider
$$\begin{array}{lcc}
A(u, x) = & \left(1 + ux + (ux)^2 + (ux)^3 + \ldots
\right) & \left( 1 + (ux^2)^1 + (ux^2)^2 + \ldots \right) \cr
& ({\rm generates \ units}) & ({\rm generates \ twos}) \cr
&&\hfill (4.6)\cr
&\left( 1 +
(ux^3)^1 + (ux^3)^2 + \ldots \right) \ \ \ \ldots \cr
&({\rm generates \ threes}) \ \ \ \
\end{array}$$
Then the terms in $u^Nx^L$ will have exactly $N$ integers in the
corresponding partition of $L$ so (cf. 4.4)
$$A(u,x) = 1 + \sum^\infty_{n=1} \sum^\infty_{\ell-1} p_1(n,\ell)
u^nx^\ell = \prod\limits^\infty_{\ell =1} (1 - ux^\ell)^{-1} \ . $$
However, we want the number of partitions with $N$ or fewer integers,
i.e., \break $p(N,L) = \sum^N_{n=1} p_1 (n,L)$. These sums will be
automatically generated if we multiply $A$ by $(1 + u + u^2 + u^3
\ldots)$ before taking the coefficient of $u^N x^L$ so
$$B_1 (u,x) = \sum^\infty_{n=1} \sum^\infty_{\ell = 0} p(n,\ell) u^n
x^\ell = (1 + u +u^2 + \ldots) A(u,x) = \prod\limits ^\infty _{\ell = 0} (1 -
ux^\ell)^{-1} \ . \eqno (4.7)$$
\vfill
\eject
Readers will recognise the analogy of this expression with the grand
partition function for a gas of non-interacting Bosons. So $p(N,L)$
can be found from the product as the coefficient of $u^Nx^L$. Had we
been interested in Bosons on a line then $p(N,L)$ would have given us
the desired function ${\overline G} (L)$ but our problem is three
dimensional. Instead of partitioning $L$ into integers $\ell_I$ we
need it partitioned into integer triples $\left( \ell_{Ix}, \ell_{Iy},
\ell_{Iz}\right) $ and the general term in our polynomial will be
$x^{\ell_{1x}}_1 y_1^{\ell_{1y}} z_1^{\ell_{1z}} x_2^{\ell_{2x}}
\ldots x_N^{\ell_{Nx}} y_N^{\ell_{Ny}} z_N^{\ell_{Nz}}$. When we
permute we do so by exchanging $(x_1, y_1, z_1)$ as a triple with say
$(x_2, y_2, z_2)$. The degree of our polynomial is
$$L= \sum^N_{I=1} \left( \ell_{Ix} + \ell _{Iy} + \ell_{Iz} \right) \
. $$
A triple $(2,1,0)$ corresponding to $\ell_{1x} = 2, \ell_{1y} = 1,
\ell_{1z} = 0$ will not be permuted into $(1,0,2)$ by exchanging
particle labels so such triples must be regarded as distinct. The
number of partitions of 3 into triples, such that the order within a
triple matters but the order of the different triples does not,
$p_3(L)$ with $L=3$, is already quite a handful. Writing $=$ for an
equal total we have
$$\left( \begin{array}{c}
3 \\
0 \\
0 \end{array}
\right) =
\left( \begin{array}{c}
0 \\
3 \\
0 \end{array}
\right)
=
\left( \begin{array}{c}
0 \\
0 \\
3 \end{array}
\right)
=
\left( \begin{array}{c}
2 \\
1 \\
0 \end{array}
\right)
=
\left( \begin{array}{c}
0 \\
2 \\
1 \end{array}
\right)
=
\left( \begin{array}{c}
1 \\
0 \\
2 \end{array}
\right)
=
\left( \begin{array}{c}
1 \\
2 \\
0 \end{array}
\right)
=
\left( \begin{array}{c}
0 \\
1 \\
2 \end{array}
\right)
=$$
$$
=
\left( \begin{array}{c}
2 \\
0 \\
1 \end{array}
\right)
=
\left( \begin{array}{c}
1 \\
1 \\
1 \end{array}
\right)
=
\left( \begin{array}{c}
2 \\
0 \\
0 \end{array}
\right)
+
\left( \begin{array}{c}
1 \\
0 \\
0 \end{array}
\right)
\ =
\begin{array}{ll}
& {\rm 8 \ more \ like \ that} \cr
&{\rm with} \ x,y,z \ {\rm components} \cr
& {\rm permuted.}
\end{array}
\ =
$$
$$
\hskip 3.4cm
\ =
\left( \begin{array}{c}
1 \\
1 \\
0 \end{array}
\right)
+
\left( \begin{array}{c}
1 \\
0 \\
0 \end{array}
\right)
\! =
\begin{array}{ll}
& {\rm 8 \ more \ like \ that} \cr
&{\rm with} \ x,y,z \ {\rm components} \cr
& {\rm permuted \ in \ each \ triple.}
\end{array}
\ =
$$
$$
\hskip 1cm
=
\left( \begin{array}{c}
1 \\
0 \\
0 \end{array}
\right)
+
\left( \begin{array}{c}
1 \\
0 \\
0 \end{array}
\right)
+
\left( \begin{array}{c}
1 \\
0 \\
0 \end{array}
\right)
=
\begin{array}{ll}
& {\rm 9 \ more \ like \ that} \cr
&{\rm (not \ 26 \ because \ the}\cr
& {\rm ordering \ of \ the \ triples} \cr
&{\rm does \ not \ count).}
\end{array}
$$
Hence there are 38 partitions of 3 into triples! Now let $p_3 (L)$ be
the number of partitions of $L$ into triples, every triple being
counted as different but the different orderings of the same triples
being regarded as the same. By analogy with (4.5) we consider the
expression
\eject
$$
A(x,y,z)\equiv $$
$${\setlength{\arraycolsep}{.1em}
\begin{array}{lllll}
\equiv (1 + x + x^2 ...)& \times & (1 +
(x^2)^1 + (x^2)^2 + ...)& \times & (1 +
(x^3)^1 + (x^3)^2 + ... )\times ... \cr
\times \, (1 + y + y^2 ...) & \times & (1 +
(y^2)^1 + (y^2)^2 + ...)& \times & (1 +
(y^3)^1 + (y^3)^2 + ... ) \cr
\times \, (1 + z+ z^2 ...) & \times & (1 +
(z^2)^1 + (z^2)^2 + ...) & \times & (1 +
(z^3)^1 + (z^3)^2 + ... ) \cr
& \times & (1 + (yz)^1 + (yz)^2 + ...) &
\times & (1 + (x^2y)^1 + (x^2y)^2 ...)\cr
{\rm This \ column} & \times & (1 + (zx)^1 +
(zx)^2 + ...) & \times & (1+ (y^2z)^1 +
(y^2z)^2 + ...) \cr
{\rm generates} & \times & (1+(xy)^1 +(xy)^2 +
...) & \times & (1+ (z^2x) + ...)\cr
{\rm triples} & & {\rm column \ generates} & \times &
(1 + (x^2z) + ...)\cr
\left( \! \begin{array}{c}
1 \\
0 \\
0 \end{array}
\! \right)
\! {\rm or} \!
\left(\! \begin{array}{c}
0 \\
1 \\
0 \end{array} \! \right)
\! {\rm or} \!
\left(\! \begin{array}{c}
0 \\
0 \\
1 \end{array}\! \right)&&
\hskip -.1cm
\left(\! \begin{array}{c}
2 \\
0 \\
0 \end{array}
\! \right) \! \!
\left( \! \begin{array}{c}
0 \\
2 \\
0 \end{array} \! \right) \! \!
\left( \! \begin{array}{c}
0 \\
0 \\
2 \end{array}\! \right) \! \!
\left( \! \begin{array}{c}
1 \\
1 \\
0 \end{array}
\! \right) \! \!
\left( \! \begin{array}{c}
0 \\
1 \\
1 \end{array}\! \right) \! \!
\left( \! \begin{array}{c}
1 \\
0 \\
1 \end{array} \! \right)&&
\hskip -0.45cm
\begin{array}{rl}
\times &(1 + (y^2x) + ...)\cr
\times & (1 + (z^2y) + ...)\cr
\times & (1 + (xyz) + (xyz)^2 + ...)
\end{array}\cr
{\rm and \ multiple}&& {\rm and \ multiple} && {\rm generates \
triples} \cr
{\rm combinations} && {\rm combinations} && {\rm with}\ \ell_x +
\ell_y + \ell_z =3 \cr
{\rm thereof} && {\rm thereof} &&
\end{array}}
$$
By analogy with the arguments beneath (4.5) one sees how the terms of
the third degree generate all the partitions of 3 into triples that we
have just enumerated. So the coefficient of $t^L$ in $A(t,t,t)$ will
be $p_3(L)$, the number of partitions of $L$ into triples.
Furthermore $A(x,y,z)$ may be compactly written in the form
$${\setlength{\arraycolsep}{.1em}
\begin{array}{lcl}
A(x,y,z) = & \prod\limits^\infty_{p=0}\prod\limits^\infty_{q=0}
\prod\limits^\infty_{r=0} & (1-x^py^qz^r)^{-1} \ . \cr
& ^{p + q + r \neq 0} &\cr
\end{array}}
$$
\vskip -1.7cm
$$\eqno{(4.8)}$$
\vskip .8cm
For our $N$ boson problem we are interested not in such partitions
into any triples but in partitions constrained to have $N$ or fewer
triples as summands. To get exactly $N$ summands we merely insert a
$u$ in each term as was done in (4.6). While to get the sum of all
terms with $N$ or less summands we have to multiply by $(1+u+u^2
\ldots)$ as in (4.7). Thus putting $x=y=z=t$ the Boson generating
function in 3 dimensions is, writing $\ell = p+q+r$,
$$B(u,t) = \prod\limits^\infty_{p=0} \prod\limits^\infty _{q=0}
\prod\limits^\infty _{r=0} \left(1 - ut^{p+q+r} \right)^{-1} =
\prod\limits^\infty_{\ell = 0} \left(1 - ut^\ell \right)^{-{1 \over 2} (\ell
+1)(\ell +2)} \, \eqno (4.9)$$
where ${1 \over 2} (\ell +1)(\ell +2)$ is the number of terms in the
triple product with $p+q+r = \ell$.
The required function ${\overline G} (L)$ is the coefficient of $u^N
t^L$ in $B(u,t)$. Again if one only wants ${\overline G} (L)$ for
$L<L_{\max}$ then the infinite product can be replaced by a finite
product up to $L_{\max}$ without altering the required coefficients.
To get ${\overline G}_1(L)$ one merely takes the coefficient of
$u^Nt^L$ in $(1 - t)^3B(u,t)$ \break while ${\overline G}_1 (L) - {\overline
G}_1 (L-2)$ is the coefficient of $u^Nt^L$ in $(1 -t^2) (1 -t)^3
B(u,t) = (1+t) (1-t)^4 B(u,t)$.
In summary the degeneracy $g_B(L)$ of the $N$ Boson state with
hyperangular momentum $L$ is the coefficient of $u^N t^L$ in the
expression
$$(1+t) (1-t)^4 \prod\limits^\infty_{n=0} \left( 1 -
ut^{\ell}\right)^{-{1 \over 2} \left(\ell + 1 \right) \left( \ell +2
\right)} \ \ . \eqno (4.10)$$
For two particles $N=2$ we find that the coefficient of $u^2$ is
$\left( 1-3t^2 \right) \left( 1 - t^2 \right)^{-2}$. The coefficient
of $t^L$ in this expression is $2L+1$ when $L$ is even and zero when
$L$ is odd just as it should be. This is just as in the $C^{12} - C^{12}$
homonuclear diatomic molecule of two Bosons with every odd rotational state
missing as seen in Carbon star spectra.
\subsection{Fermions}
The argument of \sect3 relates the number of hyperspherical
harmonics of degree $L$ to the number of harmonic polynomials of that
degree. Our experience in \sect4$\, a$ leads us to study first the
number of antisymmetric polynomials of degree $L$ in $N$ dimensions.
If $x^{\ell_1}_1 x^{\ell_2}_2 \ldots x^{\ell_N}_N$ is a term in such a
polynomial then $\Sigma \ell_I = L$. Furthermore the
antisymmetric polynomial involving that term is the Slater determinant
$$\left| \begin{array}{ccccl}
x^{\ell_1}_1 & x_2^{\ell_1} & \ldots & x^{\ell_1}_N \cr
x^{\ell_2}_1 & x^{\ell_2}_2 & \ldots & x^{\ell_2}_N \cr
\vdots &&&\vdots \cr
x^{\ell_N}_1 & x^{\ell_N}_2 & \ldots & x^{\ell_N}_N
\end{array}
\right|
$$
Clearly if $\ell_I = \ell_J$ for $I \neq J$ then this determinant
vanishes. Furthermore the determinant only changes sign (at most) if
the $\ell_I$ are permuted. So for any term that survives we may,
without loss of generality, take $\ell_1 < \ell_2 < \ell_3 \ldots <
\ell_N$. Thus among the $\ell_I$ only $\ell_1$ can be zero, $\ell_2$ must
be at least 1, $\ell_3$ at least 2 and so on with $\ell_N$ at least $N-1$.
For a non-zero result $L= \Sigma \ell_I \geq {\textstyle {1 \over
2}} (N-1)N$.
By analogy with our study of partitions $p (L)$ for the Boson case we
now study partitions of $L$ into distinct parts. Let $q(L)$ be the number
of decompositions of $L$ into distinct integer summands without regard
to order. Thus $5 = 4+1 = 3+2$ so that $q(5) = 3$. The generating
functions for the $q(L)$ is, setting $q(0)=1$,
$$\sum^\infty _0 q(L) x^L = \prod\limits^\infty_{\ell = 1} (1 + x^\ell) \
. \eqno (4.11)$$
However not all decompositions of $L$ into distinct parts lead to
antisymmetric polynomials in $N$ dimensions. We need a decomposition
into either $0 + (N-1)$ unequal integers or into $N$ unequal integers,
i.e., we need the coefficient of $u^N x^L$ in
$$\prod\limits^\infty_{\ell =0} \left( 1 + ux^\ell \right) =
\sum^\infty_{L=0} \sum^\infty _{N=0} q (L,N) u^N x^L \ , \eqno
(4.12)$$
which is the expression analogous to (4.7) of the Boson case. Again
it is the analogue of the grand partition function for Fermions.
The generalisation corresponding to three dimensions follows the
argument for (4.8) and yields the generating function
$${\setlength{\arraycolsep}{.1em}
\begin{array}{lcl}
E(x,y,z) = & \prod\limits^\infty_{p=0}\prod\limits^\infty_{q=0}
\prod\limits^\infty_{r=0} & (1+x^py^qz^r) \ , \cr
& ^{p + q + r \neq 0} &\cr
\end{array}}
$$
\vskip -1.8cm
$$\eqno{(4.13)}$$
\vskip .7cm
\noindent
which leads analogously to (4.9) to the Fermion generating function
$$F(u,t) = \prod\limits^\infty_{p=0} \prod\limits^\infty_{q=0}
\prod\limits^\infty_{r=0} \, \left( 1 + u t^{p+q+r} \right) =
\prod\limits^\infty _{\ell = 0} \left( 1 + ut^{\ell}\right) ^{{1 \over
2} (\ell +1)(\ell +2)} \ . \eqno (4.14)$$
The coefficient of $u^Nt^L$ in this expression gives the number of
homogeneous polynomials of degree $L$ antisymmetric for interchanges
of triples in 3$N$ dimensions. For Fermions of spin ${\textstyle {1
\over 2}}$ we do not need complete antisymmetry but only antisymmetry
between particles of the same spin state. To allow for the $\alpha$
and $\beta$ spin states being alternatives we generalise $E (x,y,z)$
to
$$\prod\limits^\infty_{p=0} \prod\limits^\infty_{q=0}
\prod\limits^\infty_{r=0} \, \left( 1 + x^py^qz^r \alpha \right)
\left( 1 + x^py^q z^r \beta \right) \eqno (4.15)$$
and in place of $F(u,t)$ we then find $F^2$
$$F^2 = \prod\limits^\infty_{\ell = 0} \left( 1 + u t^\ell
\right)^{(\ell + 1)(\ell+2)} \ . \eqno (4.16)$$
The coefficient of $u^Nt^L$ in $F^2$ is the number of Fermionic
polynomials for spin ${1 \over 2}$ particles with the correct
antisymmetry. The arguments relating the degeneracy of the
$N$-Fermion spin ${1 \over 2}$ wave function for the state of
hyper-angular momentum $L$ to this expression is the same as for the
Boson case. Thus $g_F(L)$ is coefficient of $u^Nt^L$ in the expression
$$(1+t) (1-t)^4 \prod\limits^\infty_{\ell = 0} \left( 1 + ut^\ell
\right)^{(\ell +1) (\ell +2)} \ . \eqno (4.17)$$
To check this formula we take the coefficient of $u^2$, and find
for the 2 Fermion problem the generating function
$$\left( 1 + 9t + 3t^2 + 3t^3 \right) \left( 1 - t^2 \right) ^{-2} =
\sum_{\ell {\rm even}} (2 \ell + 1 ) t^\ell + 3 \sum_{\ell {\rm odd}} (2
\ell + 1 ) t^{\ell} \ , \eqno (4.18)$$
in which the coefficients of $t^\ell$ will be recognised as the
degeneracies of the rotational states of the hydrogen molecule with 2
protons of spin $1 \over 2$. The even $\ell$ values correspond to
para-hydrogen and the odd $\ell$ values to ortho-hydrogen.
We now turn to the degeneracies in hydrogen-like potentials in which
different $L$ states can be degenerate. Here we need the sum of the
$g_F(L)$ for all \break $L \leq n-1$ as in equation (3.3). If we multiply
our generating function for $g_F(L)$ by $1+t+t^2+t^3+ \ldots =
1/(1-t)$ and then pick the coefficient
of $u^Nt^{n-1}$ we will get the required sum so the hydrogenic Fermi
degeneracy is that coefficient in $$(1+t) (1-t)^3
\prod\limits^\infty_{\ell = 0} (1 + ut^\ell)^{(\ell + 1)(\ell + 2)}\
. \eqno (4.19)$$
To find the ground state we need the first energy level $n$ for which
the coefficient of $u^Nt^{n-1}$ is non-zero. To get $u^N$ and no more
with the lowest power of $t$, we need to use the $ut^\ell$ terms
rather than the 1 in all the low $\ell$ brackets since others come
with higher powers of $t$. Thus if the highest $\ell$ needed is
$\ell_m$ we require
$$\sum ^{\ell_m}_{\ell = 0} (\ell +1) (\ell +2) = N\ . \eqno (4.20)$$ The sum is
$\, {1 \over 3} (\ell _m +1)\, (\ell_m +2)\, (\ell _m + 3)\, $ but $\, N\, $ may not
be of precisely this form for integer $\ell_m$, in which case we take
the lowest $\ell_m$ that gives \break ${1 \over 3} (\ell _m +1) (\ell_m +2)
(\ell _m + 3) \geq N$ so that there will be at least one term in
$u^N$. We are interested in the least power of $t$ associated with
this term. This will be $$n-1 = \sum^{\ell_m}_{\ell =1} \ell (\ell +1)
(\ell +2) = {{\textstyle {1 \over 4}}} \ell _m (\ell _m+1) (\ell_m +2)
(\ell _m + 3) \eqno (4.21)$$ whenever $N$ is of the special closed shell form given
by the equality. Then \break $\ell_m \simeq (3N)^{1/3} - 2 - (3N)^{-1/3}$
and $n = {3N \over 4} \ell_m + 1 \rightarrow (3N)^{4/3}\big / 4$ as
$N \rightarrow \infty$. For large $N$ we then find that the ground
state energy behaves as
$$E = - {\hbar^2 \over 8M} \, {\zeta^2 \over \left[ {(3N) \over 4}
^{4/3} \right]^2} \ \, . \eqno (4.22) $$ To compare this energy with that of a white dwarf
star we must first choose an appropriate value of $\zeta$ so that the
potential corresponds to gravity. We showed in an earlier paper
(Lynden-Bell \& Lynden-Bell 1999) that
at high temperatures our systems have a Gaussian density profile at
equilibrium. For such a profile the potential energy may be expressed
in terms of the total mass $Nm_H$ and the rms radius at equilibrium
$r$ and is $$V = \left( {3 \over 4 \pi}\right)^{1/2} G (Nm_H)^2/r\
. \eqno (4.23) $$ We therefore choose $${\hbar ^2 \over 2 M} \zeta = \left( {3
\over 4 \pi} \right) ^{1/2} G (Nm_H)^2 \ . \eqno (4.24)$$ With this choice our
ground state energy level is $$E = - {2 \ G^2m^4_H m_e \over 3^{5/3}\
\pi \ \hbar ^2} N^{7/3} \eqno (4.25)$$ where we have written $M = Nm_e$ for the
mass of the degenerate electrons whose wave function we have been
evaluating. This expression is of precisely the form found by the
standard equation of state of a cold degenerate non-relativistic gas
under its own gravity. We have, therefore, established that White
Dwarfs may be regarded as `superatoms' -- systems with N-body wave
functions which are solutions of Schr\"odinger's equation in a central
potential.
\section{Conclusions}
By treating the N-body problem as a separable system in $3N$
dimensions we have shown that it can be solved in appropriate
potentials. Marshall \& Wojciechowski (1988) have given the general
form of potentials that allow separability in many dimensions and we
have concentrated on the hyper-spherical one. Of the many others
possible, a sub-class are symmetrical for exchange of the particles.
Whereas we have shown how to classify the wave functions by
hyper-angular momentum, those interested in rotating systems will need
to develop methods of classification involving the 3 dimensional
angular momentum; here the methods of Dragt (1965) and the work by
Louck \& Galbraith (1972) may prove useful. It is hoped that study of
that problem will throw light on Bose-Einstein condensation of small
clusters in rotating systems.
The fact that we are only able to treat non-relativistically
degenerate white dwarfs suggests that an appropriate generalisation
for relativistically moving particles should be sought for systems
that do not radiate gravitational waves.
In our earlier paper on the classical mechanics of these systems we
showed that the fundamental breathing oscillation in $r$ separates
for the far-more-general potentials $V= V_0 (r) + r^{-2} V_2 ({\hat
{\bf r}})$ where $V_2$ is an arbitrary function of all the coordinates
which is independent of scale, $\lambda$, when all the $x_I
\rightarrow \lambda x_I$. While this is still the case in quantum
mechanics the energy eigenvalues depend on the other motions so this
separation does not by itself yield eigenvalues. However, the
possibility of solid-like and liquid-like states where $V_0 = {1 \over
2} kr^2$ and \break $r^{-2} V_2 = {\sum \atop I} {\sum \atop {\! \! \! \! \!
<J}} k' \left| {\bf r}_I - {\bf r}_J \right|^{-2}$ suggests that such
systems are worthy of further study.
The statistical mechanics of the systems with $V = -{GM^2 \over r} \,
, r<r_e$, gives negative specific heats just as those studied earlier
as examples of phase transitions (Lynden-Bell \& Lynden-Bell 1977).
However, within one system there is no gravothermal catastrophe
(Antonov 1962, Lynden-Bell \& Wood 1968).
\begin{acknowledgments}
We thank Dr Dragt for sending us his papers relevant to classifying
our states by symmetry and Drs Twambey and Sridhar for referring us
to the Calogero model. D. Lynden-Bell is currently supported on a
PPARC Senior Research Fellowship.
We thank the referees for detailed comments that broadened the background of the paper and improved its clarity.
\end{acknowledgments}
|
\section{Introduction}
The transient \xray\ source \srcnm\ is an accreting \xray\ pulsar in an
eccentric 24 day orbit (\cite{bil97}) with the O9e star,
V635 Cassiopeiae (\cite{ung98}). \xray\ outbursts have been
observed from \srcnm\ with
\emph{Uhuru} (\cite{for76}), {\em HEAO-1}\ (\cite{whe79,ros79}),
{\it Ginga}\ (e.g. \cite{tam92}), and {\it CGRO}/BATSE (\cite{bil97}).
A cyclotron resonance scattering feature (CRSF) in \srcnm\ was first
noted near 20\,keV by Wheaton, \etal\ (1979) with the UCSD/MIT hard \xray\
(10-100\,keV) experiment aboard {\em HEAO-1}. White, Swank \& Holt (1983) analyzed
concurrent data from the lower energy (2-50\,keV) {\em HEAO-1}/A2
experiment and found an additional feature at \aprx12\,keV. Two
outbursts of \srcnm\ were observed with {\it Ginga}, in 1990 February and
1991 April (\cite{nag91,tam92,mih95}). The pattern of absorption
features differed dramatically in the two outbursts. A pair of
features similar to the {\em HEAO-1}\ results in the 1990 outburst gave
way to a single feature near 17\,keV in 1991 (\cite{mih98}).
We discuss here spectral and timing analyses of observations
of the 1999 March outburst (\cite{wil99,hei99}) obtained with the
\emph{Rossi X-Ray Timing Explorer} ({\em RXTE}).
\section{Observations and Analysis}
Observations were made with the Proportional
Counter Array (PCA) (\cite{jah96}) and High Energy X-ray Timing
Experiment (HEXTE) (\cite{rot98}) on board {\em RXTE}. The PCA is a set of 5
Xenon proportional counters sensitive in the
energy range 2--60\,keV with a total effective area of \aprx
7000\,$\rm cm^2$. HEXTE\ consists of two arrays of 4 NaI(Tl)/CsI(Na)
phoswich scintillation counters (15-250\,keV) totaling \aprx 1600
{$\rm cm^2$}. The HEXTE\ alternates between source and background
fields in order to measure the background. The PCA and HEXTE\ fields
of view are collimated to the same 1\hbox{$^\circ$}\ full width half
maximum (FWHM) region.
Beginning on 1999 March 3, daily, short (\aprx 1\,ks) monitoring
observations were carried out. In addition, we performed four long
pointings (duration \aprx15-35\,ks, labeled A -- D in
Fig.~\ref{f_asm}) to search for CRSFs.
Observation B, on 1999 March 11.87-12.32, spanned
periastron passage at March 11.95 (\cite{bil97}). Figure~\ref{f_asm}
shows the RXTE/All Sky Monitor (ASM, 1.5--12\,keV) light curve of
\srcnm\ together with the times of the pointed observations. In this
work, we concentrate on observation B.
\begin{figure}
\centerline{\includegraphics[angle=0,width=3.5in]{0115asm_rev.eps}}
\figcaption{\label{f_asm} The {\em RXTE}/ASM light curve of \srcnm\ averaged in
6\,h segments. The vertical dotted lines indicate times of periastron
passage. The bars at the top indicate the times of the short,
public {\em RXTE}\ pointings and of the long observations (heavy bars, A--D).}
\end{figure}
\subsection{\label{s_anal}Spectral Analysis}
The spectrum of \srcnm\ varies significantly with neutron star
rotation phase (\cite{nag91}), making fits to the average spectrum
difficult to interpret. In order to study the evolution of the
spectrum through the pulse, we corrected photon arrival times to both
the solar system and the binary system barycenters using the ephemeris
of Bildsten, \etal\ (1997). We then applied a Z$^2$ period search
(\cite{buc83}) to the HEXTE\ data to determine the pulse period.
Figure~\ref{f_flc} shows folded light curves derived from spectra in
50 pulse phase bins for observation B, where the period was
3.614512(33)s. The folded light curve has a sharp main peak, followed
by a broader, softer second peak, similar to earlier reports
(\cite{wsh83,bil97}). In searching for CRSFs, we followed the spectral
analysis methods described in Kreykenbohm, \etal\ (1998).
\begin{figure}
\centerline{\includegraphics[angle=0,width=4.25in]{flc_rev.ps}}
\figcaption{\label{f_flc} Top: The {\em RXTE}\ folded light curve of
\srcnm\ in 5 energy bands. Bottom: F-Test probability for the addition
of a second and subsequently a third CRSF to a simple continuum fit to
the HEXTE data in 20 phase bins. In both panels, points which fall on
the dashed line at $\rm 10^0$ indicate no improvement of the fit for
addition of that line. Vertical dotted lines delineate pulse
phase 0.70--0.76 corresponding to the spectrum in Fig.~\ref{f_spec}.}
\end{figure}
Because CRSFs at \aprx12 and \aprx22\,keV are known in \srcnm\
(\cite{wsh83,nag91}), we first concentrated on the HEXTE\ data where
higher harmonics might appear. We fit a ``cut-off power law'' (a
power law times an exponential) to the HEXTE\ spectra in 20 phase bins.
The reduced \chisq\ of these fits ranged from 1.3 to 12.3 (62 degrees
of freedom, ``dof''). Concentrated at the main pulse and through the
rise and peak of the second, there were significant residuals
resembling absorption features near 20--25\,keV. In the fall of the
second peak, residuals appeared between 30--40\,keV. We then fit for
an absorption feature near 20--25\,keV, resulting in reduced \chisq s
between 0.7 and 2.0 (59 dof). We used a simple Gaussian model for the
optical depth profile. Fig.~\ref{f_flc}, shows the result of an F-Test
for adding this line. In the cases where no line was allowed by the
fit, the points are plotted with a value of $\rm 10^0$. Next, we
allowed a CRSF between 28--45\,keV. This significantly improved the
fits in about half of the phase bins, including some near the main
peak where large \aprx20\,keV residuals in the initial fits masked the
presence of this line. The corresponding F-test results are also
plotted in Fig.~\ref{f_flc}. Although there is strong evidence for
multiple lines at other phases, the phase range 0.70--0.76 shared both
a significant \aprx20\,keV line as well as the most clearly line-like
residuals in 30--40\,keV in the no-lines fit. We therefore chose to
concentrate on this phase in this \emph{Letter}. We plan to perform
detailed analysis of all phases and all four long (A--D) observations
in a future paper.
Next, we jointly fit the HEXTE and PCA data for phase 0.70--0.76. To
account for uncertainties in the response matrix, 1\% systematic
errors were applied to the PCA data. None of the continuum models
(high energy cut-off power law, Fermi-Dirac cut-off (FDCO) times a power law,
and Negative and Positive power law Exponential (NPEX); see
Kreykenbohm, \etal\ 1998) typically used for accreting pulsar spectra
provided an acceptable fit without the inclusion of absorption
features. A black body with
kT\aprx0.4\,keV and photoelectric absorption were required to describe
the data. No Fe-K line was required. Ultimately, it was necessary to
include CRSFs at \aprx12, \aprx21, and \aprx34\,keV in the joint
spectrum. The fitted line parameters were insensitive to the details
of the continuum model used. The results given here used a
Fermi-Dirac cut-off (FDCO) times a power law, given by $F(E) \propto
(1 + e^{(E - E_c)/E_f})^{-1} \times E^{-\Gamma}$. F is the photon
flux, $\rm E_c$ the cutoff energy, $\rm E_f$ the folding energy, and
$\Gamma$ the photon index. $\rm E_c$ was fixed at zero. An F-Test for
the significance of adding the \aprx34\,keV line to a model with only
two absorption features gave a chance probability of $\rm 10^{-17}$.
\subsection{Temporal Variability}
Along with standard Fourier techniques, we analyzed the data in the
time domain using the linear state space model (LSSM) formalism
described by K{\"o}nig \& Timmer (1997) and Pottschmidt, \etal\ (1998).
Parameters of the LSSM are related to
dynamical timescales of the system such as oscillation periods, decay
times of damped oscillators and stochastic noise. As shown in
Figure~\ref{f_lc}, the data are well described by a LSSM of order
2. This model, based on an auto-regressive process, is
dominated by a stochastically driven sinusoid of period 552\,s which
exponentially decays with an e-folding time of $P_{\rm fold} =
282$\,s. This short $P_{\rm fold}$ accounts for the broad QPO peak
seen in the PSD (Fig.~\ref{f_psd}). A Kolmogoroff-Smirnoff test shows
that the difference between the data and the LSSM is purely
attributable to white noise. The light curve is thus
consistent with a single exponentially decaying sinusoid, driven by
a white noise process.
\section{Results and Discussion}
\subsection{Spectrum and CRSFs}
Fig.~\ref{f_spec} shows the best fit joint count spectrum from pulse
phase 0.70--0.76. Also plotted is the inferred incident photon
spectrum. Best fit parameters are given in Table~\ref{t_fit}, and the
reduced \chisq\ of the fit is 1.66 (74 dof). This is the first time
that a fundamental and two harmonic CRSFs have been detected in a
single accreting \xray\ pulsar spectrum. Previously, at most a
fundamental and second harmonic have been seen: 4U~1907+09
(\cite{cus98}); or suggested: Vela~X-1 (\cite{kre98}), A0535+25
(\cite{ken94}), 1E~2259+586 (\cite{iwa92}). Those earlier
observations lacked the broad-band sensitivity of {\em RXTE}. Furthermore,
simple fits to the phase-resolved HEXTE\ spectra show that the \xray\
spectrum varies rapidly with neutron star rotation. We observe
significant variations between consecutive phase bins which cover only
2\% of the pulse phase. This suggests complex spatial variations of
conditions near the neutron star polar cap. Since the
35\,keV CRSF is only present in about half of the pulse (predominantly
during the fall of the second, weaker pulse), and both the 22 and
35\,keV lines are only present together in 3 of 20 coarse phase bins,
averaging over large phase angles would likely have washed out the
line in the variable continuum.
\begin{figure}
\centerline{\includegraphics[angle=0,width=3.5in]{best_fit_rev.ps}}
\figcaption{\label{f_spec} Top: The PCA and HEXTE\ count spectra (crosses) for
pulse phase 0.70-0.76. Also shown are the best fit model (histograms)
with 3 CRSFs and the inferred incident spectrum (smooth curve).
Middle: The ratio of the data to the best fit model. Bottom: The
ratio of the data to a model fit with only two CRSFs. The
residuals between 30--40\,keV and the underprediction of the continuum
above 40\,keV emphasize the presence of the third line. }
\end{figure}
The fundamental energy of 12.4\,keV implies a neutron star surface
field of $\rm 1.1\times 10^{12}(1 + z)$\,G. Contrary to simple Landau
theory, the observed line spacing is not quite harmonic.
The ratio of the $\rm 2^{nd}$ and $\rm 3^{rd}$ harmonics to the
fundamental are $\rm 1.73 \pm 0.08$ and $\rm 2.71 \pm 0.13$
respectively. We tried fits with the $\rm
2^{nd}$ and $\rm 3^{rd}$ lines constrained to be exact
harmonics of the first, whose energy was allowed to vary. The
resulting fit was unacceptable (reduced \chisq\ of 6.7 with 76
dof). We did, however, find a reasonable fit with the
$\rm 3^{rd}$ harmonic tied to the first but the second free to
vary. Nevertheless, an F-test comparing this fit to our best model gives a chance
improvement probability of $\rm 2 \times 10^{-4}$ for allowing the
non-integer energy ratio of the first and third lines.
\begin{table}
\caption{\label{t_fit} Parameters of the best fit model spectrum (see
\S \ref{s_anal}).}
\begin{minipage}{\linewidth}
\renewcommand{\thefootnote}{\thempfootnote}
\begin{tabular}{lccc} \hline \hline
Harmonic & Energy & Width\footnote{$\sigma$ of the Gaussian optical
depth profile} & Optical Depth \\
& (keV) & (keV) & \\ \hline
1 (Fundamental) & $\rm 12.40^{+0.65}_{-0.35}$ & $\rm 3.3^{+0.1.9}_{-0.4} $
& $\rm 0.72^{+0.10}_{-0.17} $\\
2 & $\rm 21.45^{+0.25}_{-0.38} $ &$\rm 4.5^{+0.7}_{-0.9} $
& $\rm 1.24^{+0.04}_{-0.06} $\\
3 & $\rm 33.56^{+0.70}_{-0.90} $ & $\rm 3.8^{+1.5}_{-0.9}$
& $\rm 1.01^{+0.13}_{-0.14} $\\ \hline
\end{tabular}
\end{minipage}
\end{table}
In addition to relativistic shifts, line energies may deviate from
harmonic for a number of reasons. The main scattering for the
harmonics may take place in regions of different magnetic field,
either resulting from optical depth effects in the
mound of accreting matter, or for lines primarily produced at opposite magnetic
poles. It is interesting to note (see Fig.~\ref{f_flc}) that the
second and third harmonics are most significant in the main
and secondary pulses, respectively, possibly indicating origins at
opposite poles.
In our initial fits to the HEXTE\ data alone, we observed that the
20\,keV CRSF varied in both strength and energy (by 20\%) through the
pulse phase. This was first observed with {\it Ginga}\ by Nagase, \etal\
(1991) in the 1990 Feb. outburst. The line is strongest and the energy
highest (\aprx24\,keV), on the falling edge of the main pulse, which
is similar to the behavior of the CRSFs in Cen~X-3 and 4U~1626-67
(\cite{hei99b}). The {\it Ginga}\ data showed significant \aprx11 and
\aprx22\,keV lines at all 8 pulse phases analyzed (\cite{mih95}). The
HEXTE data find that the \aprx20\,keV line is not required just before
the rise of the main pulse (Fig.~\ref{f_flc}). However, with the
addition of the PCA data, which constrain the continuum and
fundamental line, it is possible that this line will be required at
all phases. In any case, strong long term variability
of the lines is known (\cite{mih95}), so differences between these
results and earlier observations are not surprising.
\subsection{Temporal Variability: a 2\,mHz QPO}
Figures~\ref{f_lc} and~\ref{f_psd} show the PCA light curve of
observation~B and its power spectral density (PSD), respectively.
Strong variability with an \aprx500\,s period is obvious. At
frequencies above 5\,mHz, the PSD can be described by an overall power law
$\propto f^{-1}$ plus peaks at the neutron star rotational frequency
and its multiples. Some accreting pulsars have shown a QPO at the beat
frequency between the neutron star rotation and the Keplerian orbit at
the inner edge of the accretion disk (\cite{fin96}). Using the
relations in Finger, Wilson \& Harmon (1996) with a surface B field strength
of $\rm 1.3 \times 10^{12}$\,G,
a distance of 6\,kpc (\cite{neg99}), and a total flux of
$\rm 2.1 \times 10^{-8}$\,\eflux, the expected beat QPO
frequency is 0.8\,Hz. Overtones of the rotational frequency confuse the
search for QPOs in this region (see Fig.~\ref{f_psd}), and
no obvious peaks apart from the pulsation were seen in the 1\,ks
short pointings or in observation B.
\begin{figure}
\centerline{\includegraphics[angle=0,width=3.5in]{0115long.eps}}
\figcaption{\label{f_lc} PCA (1.5-60\,keV) light curve in 16\,s bins of
observation B together with a second order linear state space model of
the variability. The bottom panel shows the residuals of the LSSM. }
\end{figure}
Below 5\,mHz, the PSD is dominated by a broad QPO feature from the
500\,s oscillation. The shape of this feature is complex and
asymmetric, it can neither be described by a Lorentzian line nor by
the superposition of two Lorentzian lines. The feature peaks at
$\sim$1.5\,mHz and has a FWHM of 1\,mHz ($Q=f_0/\Delta f \approx
1.5$). The excess power of the QPO with respect to the underlying red
noise component in the range from 1\,mHz to 4\,mHz is \aprx5\%
RMS. The \aprx500\,s period of this oscillation is longer than any
\xray\ QPO previously reported in an accreting \xray\ pulsar. Soong
\& Swank (1989) reported a broad 0.062\,Hz QPO in {\em HEAO-1}\ observations of a
flaring state in \srcnm\ that also did not fit into the beat
frequency model.
The QPO was probably present in several of the short (\aprx1\,ks)
observations as well, as there was apparent, slow variability on a
several hundred second timescale. In observation A, the
QPO was at most present weakly, as no clear peak is evident in the
PSD. Two possible explanations for the QPO are: 1) modulation of the
accretion flow, and 2) occultation of the beam by
intervening matter in an accretion disk. It seems unlikely that the
variability is due to modulation of the accretion flow itself. The
timescales at the neutron star pole (milliseconds) and the inner edge
of the disk (seconds) are too fast. Furthermore, the rotation period
(days) of V635Cas is too long.
We compared PCA spectra at minima of the
500\,s cycle to the spectra of the following maxima. The spectral
shape is unchanged from 2.5 -- 5\,keV, and only \aprx20\% deviations
appear at higher energies.
Because the spectrum below 5\,keV is steady through the QPO and the
flux varies by a factor of two from peak to minimum, the QPO mechanism
cannot be absorption in \emph{cold} material. If it were,
the low energy spectrum would be highly modified by photoelectric
absorption. It is possible that Thomson scattering in ionized
matter causes the variability. We suggest two possible mechanisms,
which both require that the accretion disk be viewed nearly edge
on. Both are consistent with a second order LSSM process. First, an
azimuthal warp propagating around the disk could cause the ionized
disk surface to intervene in the line of sight. In this
case, the 500\,s timescale is the time for the wave to circle the disk
(assuming a single-peaked warp). In the second picture, the
absorption takes place in a lump in the disk which orbits at a Kepler
period of 500\,s.
\section{Summary}
We have made two discoveries in the {\em RXTE}\ observations of the
1999 March outburst of \srcnm. The HEXTE\ data have revealed for the
first time in any pulsar a third harmonic CRSF. The line spacing
between the fundamental and second harmonic and between the second and
third harmonics are not equal, and are not multiples of the
fundamental line energy. We have also discovered the slowest (2\,mHz)
QPO yet observed from an accreting pulsar. It was most pronounced
during an observation spanning periastron passage of the neutron star
around its massive companion. Based on the timescale, amplitude, and
energy spectrum of the oscillation, it is most likely due to
obscuration of the neutron star by hot accretion disk matter. The
longest QPO previously observed had a timescale of 100\,s in SMC~X-1
(\cite{ang91}).
\begin{figure}[h]
\centerline{\includegraphics[angle=0,width=3.5in]{longpsd01_rev.eps}}
\figcaption{\label{f_psd}The PSD (normalized according to Miyamoto et
al. 1991) of the observation B PCA light curve. The PSD has been
rebinned such that $\rm df/f =$ 0.35, 0.1 and 0.05 in the
frequency bands $10^{-4}$--$10^{-3}$\,Hz, $10^{-3}$--$10^{-1}$\,Hz,
and $10^{-1}$--$10^{1}$\,Hz respectively. The \aprx2\,mHz QPO and
several harmonics of the 3.6\,s pulsation are clear above a 1/f
continuum. }
\end{figure}
\acknowledgements
We thank E. Smith, J. Swank, and C. Williams-Heikkila
for rapidly scheduling the observations and supplying
the realtime data. ASM data are provided by the {\em RXTE}/ASM teams at
MIT and at the {\em RXTE}\ SOF and GOF at NASA's GSFC. This work was
supported by NASA grants NAS5-30720 and NAG5-7339, DFG grant Sta
173/22, and a travel grant from the DAAD.
|
\section{Introduction}
Semiclassical trace formulas for both chaotic \cite{Gut67,Gut90} and regular
\cite{Ber76} systems relate quantum spectra and classical periodic orbits.
These formulas have proven to be useful in the analysis of level statistics
\cite{Ber85} and long-range correlations \cite{Win87} in the quantum spectra,
and it has even become possible to compute individual eigenenergies from
these expressions \cite{Cvi89,Sie91,Ber92,Mai97a,Mai98}.
Gutzwiller's trace formula \cite{Gut67,Gut90} and the Berry-Tabor formula
\cite{Ber76} are semiclassical approximations to the density of states but
do not provide information about experimentally measurable observables, i.e.,
matrix elements of Hermitian operators.
The trace formulas have been extended to the calculation of diagonal matrix
elements of smooth operators in Refs.\ \cite{Wil88,Eck92}.
The extended trace formulas relate the diagonal matrix elements of operators
to the periodic orbit means of the corresponding classical observables.
However, these formulas cannot be applied directly for the semiclassical
calculation of {\em products} of diagonal matrix element where the
weighted density of states cannot, in general, be written as a trace formula.
Products of diagonal matrix elements are important in several interesting
applications of semiclassical theories, e.g., for the semiclassical theory
of matrix element fluctuations \cite{Eck95}, with the variance of an operator
$\hat A$ in an eigenstate $|n\rangle$ given by
${\rm Var}_n A\equiv\langle n|\hat A^2|n\rangle-\langle n|\hat A|n\rangle^2$.
A semiclassical periodic orbit formula for products of diagonal matrix
elements is also of crucial importance for the semiclassical quantization
technique developed in Ref.\ \cite{Mai99}, where the classical information
of a set of observables is used to significantly improve the convergence
properties of periodic orbit quantization.
In this paper we investigate {\em non-trace} type formulas for the density
of states weighted with the diagonal matrix elements of two operators
$\hat A$ and $\hat B$, i.e., $\rho^{(A,B)}(E)
=\sum_n\langle n|\hat A|n\rangle\langle n|\hat B|n\rangle\delta(E-E_n)$.
From the high resolution analysis of quantum spectra in the semiclassical
regime we find strong evidence that weighting the $\delta$-functions in the
quantum mechanical density of states with the product of diagonal matrix
elements, $\langle n|\hat A|n\rangle\langle n|\hat B|n\rangle$, is equivalent
to weighting the periodic orbit contributions in the semiclassical periodic
orbit sum with the product of the periodic orbit means,
$\langle A\rangle_p\langle B\rangle_p$, of the classical observables
$A$ and $B$.
The outline of the paper is as follows.
In Sec.\ \ref{trace:sec} we first briefly review Gutzwiller's trace formula
for chaotic systems and the Berry-Tabor formula for integrable systems, and
discuss the extension of both equations to the calculation of diagonal
matrix elements.
We then apply the theories to systems with scaling properties, and introduce
the high resolution analysis (harmonic inversion) of quantum spectra as
a powerful tool to numerically verify the validity of the semiclassical
expressions.
In Sec.\ \ref{non-trace:sec} we present our results on the semiclassical
non-trace type formulas.
Strong numerical evidence for the validity of the non-trace type equations
is provided by the harmonic inversion of spectra of two different systems,
viz.\ the hydrogen atom in a magnetic field and the circle billiard.
Sec.\ \ref{conclusion} concludes with remarks on useful and important
applications, and an outlook on possible generalizations of the non-trace
type formulas.
\section{Semiclassical trace formulas}
\label{trace:sec}
\subsection{Matrix element extension of periodic orbit theory}
The quantum mechanical density of states can be written as the trace of
the Green's function,
$\rho(E)=\sum_n\delta(E-E_n)=-(1/\pi)\,{\rm Im}\,{\rm tr}\,\hat G_E^+$.
Replacing the quantum mechanical Green's function,
$\hat G_E^+=(E-\hat H+i\epsilon)^{-1}$ with its semiclassical analogue
and calculating integrals and traces in stationary phase approximation
Gutzwiller derived the fundamental equation of {\em periodic orbit theory}
\cite{Gut67,Gut90}, i.e., the density of states expressed in terms of
quantities of the periodic orbits of the classical system.
To obtain the density of states weighted with the diagonal matrix elements
of an operator $\hat A$ we start from the generalized trace formula
\begin{eqnarray}
\rho^{(A)}(E)
&=& -{1\over\pi}\, {\rm Im}\, {\rm tr} \left(\hat G_E^+\hat A\right)
\nonumber \\
&=& -{1\over\pi}\, \lim_{\epsilon\to 0} {\rm Im}\, \sum_n
{\langle n|\hat A|n\rangle\over E-E_n+i\epsilon}
\nonumber \\
&=& \sum_n \langle n|\hat A|n\rangle \delta(E-E_n) \; .
\label{rho_A_qm:eq}
\end{eqnarray}
The r.h.s.\ of Eq.\ \ref{rho_A_qm:eq} is the density of states weighted with
the diagonal matrix elements $\langle n|\hat A|n\rangle$ of the operator
$\hat A$.
The semiclassical approximation to Eq.\ \ref{rho_A_qm:eq} for a system with
$N$ degrees of freedom reads \cite{Wil88,Eck92}
\begin{eqnarray}
\label{rho_A_sc:eq}
& & \rho^{(A)}(E) = \rho^{(A)}_0(E) \\
&+& {1\over\pi\hbar}\, {\rm Re}\, \sum_p A_p
\sum_{r=1}^\infty {T_p\over\sqrt{|\det(M_p^r-I)|}}\,
e^{i[S_p(E)/\hbar-{\pi\over 2}\mu_p]r} \; , \nonumber
\end{eqnarray}
where the Weyl term
$\rho^{(A)}_0(E)=h^{-N} \int d{\bf q}d{\bf p} A({\bf q},{\bf p})
\delta(E-H({\bf q},{\bf p}))$
is a smooth function of the energy and the fluctuating part is given by the
periodic orbit sum, with $T_p$ the time period, $S_p$ the classical action,
$M_p$ the monodromy matrix, and $\mu_p$ the Maslov index of the primitive
periodic orbit $p$.
The integer $r$ is the repetition number of the orbit.
The weights $A_p$ in the periodic orbit sum (\ref{rho_A_sc:eq}) are the means
of the observable $A$ along the periodic orbit $p$, i.e.
\begin{equation}
A_p = {1\over T_p} \int_0^{T_p} A({\bf q}_p(t),{\bf p}_p(t)) dt \; .
\label{Ap_t:eq}
\end{equation}
The derivation of Eq.\ \ref{rho_A_sc:eq} requires smoothness of the
observable $A$ over regions in phase space of size $h^N$ \cite{Eck92}.
A rigorous mathematical proof of the semiclassical trace formula
(\ref{rho_A_sc:eq}) using a coherent states decomposition can be found
in \cite{Com97}.
In Refs.\ \cite{Wil88,Eck92,Com97} formulas for the semiclassical
calculation of diagonal matrix elements are obtained for chaotic systems
with isolated periodic orbits.
For {\em regular} systems the semiclassical trace formula for the density
of states has been derived by Berry and Tabor \cite{Ber76}.
For simplicity we restrict ourselves to systems with two degrees of freedom.
Assuming now that the Hamiltonian is classically integrable, one can
express it in action-angle variables $({\bf I},\varphi)$ with
$\varphi_1,\varphi_2\in [0,2\pi]$ as $H({\bf I})$.
For a given torus, $\omega_i=\partial H/\partial I_i$ $(i=1,2)$ are the
corresponding angular frequencies.
Periodic orbits are associated with tori such that the rotation number
$\alpha\equiv \omega_1/\omega_2$ is rational, i.e., $\alpha=M_1/M_2$ with
$M_1$ and $M_2$ integers.
The fluctuating part of the Berry-Tabor formula reads
\begin{eqnarray}
& & \rho_{\rm fl}(E) \nonumber \\
&=& {1\over\pi\hbar^{3/2}}\, {\rm Re}\, \sum_{\bf M}
{T_{\bf M}\over M_2^{3/2}|g_E''|^{1/2}} \,
e^{i\left(S_{\bf M}(E)/\hbar-{\pi\over 2}\eta_{\bf M}-{\pi\over 4}\right)}\; ,
\label{BT:eq}
\end{eqnarray}
with ${\bf M}=(M_1,M_2)$ specifying the periodic orbit, and $T_{\bf M}$,
$S_{\bf M}$, and $\eta_{\bf M}$ the time, action and Maslov index of the
orbit, respectively.
The function $g_E$ in (\ref{BT:eq}) is obtained by inverting the Hamiltonian,
expressed in terms of the actions $(I_1,I_2)$ of the corresponding torus,
with respect to $I_2$, viz.\ $H(I_1,I_2=g_E(I_1))=E$ \cite{Boh93}.
By analogy with Eq.\ \ref{rho_A_sc:eq} for chaotic systems the Berry-Tabor
formula (\ref{BT:eq}) can now be generalized straightforwardly to the
semiclassical calculation of diagonal matrix elements \cite{Meh99}, yielding
\begin{eqnarray}
\label{BT_A:eq}
& & \rho^{(A)}(E)= \rho^{(A)}_0(E) \\
&+& {1\over\pi\hbar^{3/2}}\, {\rm Re}\,
\sum_{\bf M} A_{\bf M} {T_{\bf M}\over M_2^{3/2}|g_E''|^{1/2}} \,
e^{i[S_{\bf M}(E)/\hbar-{\pi\over 2}\eta_{\bf M}-{\pi\over 4}]}\, ,
\nonumber
\end{eqnarray}
with
\begin{equation}
A_{\bf M} = {1\over(2\pi)^2}\int_0^{2\pi}d\varphi_1\int_0^{2\pi}d\varphi_2
A(I_1,I_2,\varphi_1,\varphi_2)
\label{Ap_reg:eq}
\end{equation}
the classical average of the observable $A$ on the torus.
\subsection{Scaling systems}
In the following we will apply Eqs.\ \ref{rho_A_sc:eq} and \ref{BT_A:eq}
to systems with scaling properties.
In scaling systems the classical phase space structure does not change
for all values of an appropriate scaling parameter, $w$.
The scaling parameter is usually some power of an external field strength
or, for Hamiltonians with homogeneous potentials, the energy.
Examples are billiard systems \cite{Meh99} or atoms in magnetic fields
\cite{Fri89,Has89,Wat93}.
In scaling systems the shape of periodic orbits does not depend on the
scaling parameter, $w$, and the classical action $S_p$ scales as
\begin{equation}
S_p = w s_p \; .
\label{S_po}
\end{equation}
The scaling parameter plays the role of an inverse effective Planck constant,
i.e., $w\equiv\hbar_{\rm eff}^{-1}$.
For scaling systems the weighted densities of states, Eqs.\ \ref{rho_A_sc:eq}
and \ref{BT_A:eq} can be rewritten as a function of the scaling parameter $w$,
i.e.
\begin{eqnarray}
&& \rho^{(A)}(w)
= \rho^{(A)}_0(w) \nonumber \\
&+& {1\over\pi\hbar}\, {\rm Re}\, \sum_p A_p \sum_{r=1}^\infty
{s_p\over\sqrt{|\det(M_p^r-I)|}}\, e^{i[s_p w-{\pi\over 2}\mu_p]r}
\label{rho_A_scaled:eq}
\end{eqnarray}
for chaotic systems, and
\begin{eqnarray}
&&\rho^{(A)}(w) = \rho^{(A)}_0(w) \nonumber \\
&+& {1\over\pi\hbar^{3/2}}\, {\rm Re}\, \sum_{\bf M} A_{\bf M}
{s_{\bf M}\over M_2^{3/2}|g_E''|^{1/2}} \,
e^{i[s_{\bf M}w-{\pi\over 2}\eta_{\bf M}-{\pi\over 4}]}
\label{BT_A_scaled:eq}
\end{eqnarray}
for two-dimensional systems with regular dynamics.
Note that the time periods $T_p$ and $T_{\bf M}$ in Eqs.\ \ref{rho_A_sc:eq}
and \ref{BT_A:eq} must be replaced with the scaled actions $s_p$ and
$s_{\bf M}$.
Furthermore the time average of the classical observable $A$
(Eq.\ \ref{Ap_t:eq}) must be replaced with the average with respect to the
scaled action,
\begin{equation}
A_p = {1\over s_p} \int_0^{s_p} A({\bf q}_p(s),{\bf p}_p(s)) ds \; .
\label{Ap_s:eq}
\end{equation}
If an observable $A$ is chosen which is invariant under the scaling of
the system [or scales $\sim w^\beta$ with a constant exponent $\beta$]
the periodic orbit amplitudes and scaled actions in Eqs.\
\ref{rho_A_scaled:eq} and \ref{BT_A_scaled:eq} do not depend on $w$
[despite a possible power law scaling of the amplitudes with $w^\beta$
which can be transfered to the l.h.s.\ of Eqs. \ref{rho_A_scaled:eq}
and \ref{BT_A_scaled:eq}].
The attractive feature of scaling systems is that the semiclassical
weighted density of states [or more generally the density of states
multiplied by $w^\beta$] is a superposition of sinusoidal functions of
the scaling parameter $w$.
The Fourier transforms of $w^\beta\rho^{(A)}(w)$ should therefore exhibit
sharp peaks at the positions of the scaled actions of the periodic orbits.
When analyzing quantum spectra, we will make use of the scaling advantages
in the following.
\subsection{Precision check of the semiclassical trace formulas}
We now wish to apply the semiclassical trace formulas, Eqs.\
\ref{rho_A_scaled:eq} and \ref{BT_A_scaled:eq}, to a physical system with
chaotic and regular dynamics, respectively, and to check numerically the
validity of the semiclassical equations.
The numerical check is not motivated by doubts on the validity of these
expressions, which have been mathematically proven, rather we want to
introduce a powerful numerical technique for the high precision check of
equations of this kind.
We will demonstrate the accuracy of the method on the well established
semiclassical trace formulas here and then apply the same technique to
numerically verify our conjecture on semiclassical non-trace type formulas
in Sec.\ \ref{non-trace:sec}.
The semiclassical trace formulas can be tested, in principle, by the
Fourier transform analysis of quantum spectra.
The Fourier transformed spectra should exhibit peaks at the periods
(scaled actions) of periodic orbits with amplitudes given by the
semiclassical expressions.
However, the transformation of spectra with finite length yields limited
resolution only, due to the uncertainty principle of the Fourier transform,
which implies a fundamental restriction to high precision checks of the
semiclassical trace formulas.
We therefore adopt the method of Ref.\ \cite{Mai97b} where we introduced
{\em harmonic inversion} as a high resolution method for the analysis of
quantum spectra.
We briefly review the basic ideas of the harmonic inversion technique
and refer the reader to Ref.\ \cite{Mai97b} for more details.
According to Eqs.\ \ref{rho_A_scaled:eq} and \ref{BT_A_scaled:eq} the
semiclassical weighted density of states can be written as the sum of a
smooth background $\rho^{(A)}_0(w)$ and oscillatory modulations induced
by the periodic orbits,
\begin{equation}
\label{semicl}
\rho^{(A)}(w) = \rho^{(A)}_0(w) + {\rm Re} \sum_p d^{(A)}_p e^{i s_p w} \; .
\end{equation}
The amplitudes $d^{(A)}_p$ and scaled actions $s_p$ of the periodic orbits
are obtained from classical calculations and are in general complex
quantities.
The amplitudes $d^{(A)}_p$ contain the phase information determined by the
Maslov indices of orbits and the classical means of the observable $A$
given by Eqs.\ \ref{Ap_reg:eq} and \ref{Ap_s:eq} for regular and chaotic
systems, respectively.
Instead of using the standard Fourier analysis to extract the amplitudes
and actions, we adjust a finite range of the quantum spectrum by the
semiclassical expression (\ref{semicl}) with unknown and in general complex
parameters $d^{(A)}_p$ and $s_p$.
The problem of fitting a ``signal'' $\rho^{(A)}(w)$ to the functional form
(\ref{semicl}) is known as harmonic inversion.
As a numerical technique for the harmonic inversion of a signal, i.e.\ a
quantum spectrum, we apply the method of filter-diagonalization
\cite{Wal95,Man97} which allows extracting the spectral quantities in any
given interval of interest.
Operationally, one proceeds by setting up a small generalized eigenvalue
problem.
The actions $s_p$ in the chosen spectral domain and amplitudes
$d^{(A)}_p$ are obtained from the resulting eigenvalues and eigenvectors.
Thus, the recurrence spectrum is effectively discretized, the number
of terms being the number of eigenvalues in the spectral domain.
This method is a variational one (as opposed to the Fourier
transform) and therefore practically has an infinite resolution
once the amount of information contained in the signal $\rho^{(A)}(w)$
is greater than the total number of unknowns $d^{(A)}_p$ and $s_p$.
As a physical system for the high precision analysis of quantum spectra
and the comparison with the semiclassical trace formulas we choose the
hydrogen atom in a magnetic field \cite{Fri89,Has89,Wat93}.
This is a scaling system, with $w=\gamma^{-1/3}=\hbar_{\rm eff}^{-1}$ the
scaling parameter and $\gamma=B/(2.35\times 10^5\, {\rm T})$ the magnetic
field strength in atomic units.
Introducing scaled coordinates $\gamma^{2/3}{\bf r}$ and momenta
$\gamma^{-1/3}{\bf p}$ and choosing the projection of the angular momentum
on the magnetic field axis $L_z=0$ one arrives at the scaled Hamiltonian
\begin{equation}
\tilde H = {1\over 2}{\bf p}^2 - {1\over r} + {1\over 8}(x^2+y^2)
= \tilde E \; ,
\label{H_scal:eq}
\end{equation}
with $\tilde E=E\gamma^{-2/3}$ the scaled energy.
The classical dynamics is near-integrable at low energies, $\tilde E<-0.5$,
and undergoes a transition from regularity to chaos in the energy region
$-0.5<\tilde E<-0.13$.
At energies above $\tilde E=-0.13$ a Poincar\'e surface of section analysis
of the classical dynamics does not exhibit any regular structures larger
than of microscopic size \cite{Has89}.
We compare spectra at constant scaled energy $\tilde E=-0.1$ with the
results of the semiclassical trace formula (\ref{rho_A_scaled:eq}) for
chaotic systems, and spectra in the near-integrable regime at $\tilde E=-0.5$
with the extended Berry-Tabor formula (\ref{BT_A_scaled:eq}).
We choose two different operators.
The first,
\begin{equation}
\hat A = {1\over 2r{\bf p}^2}
\label{Adef:eq}
\end{equation}
has already served to study the distribution of transition matrix elements
in classically chaotic and mixed quantum systems \cite{Boo95,Boo96}.
The second operator is
\begin{equation}
\hat B = r{\bf p}^2 \; .
\label{Bdef:eq}
\end{equation}
Eigenvalues of the scaling parameter $w$ are obtained by solving
Schr\"odinger's equation (in semiparabolic coordinates
$\mu=\sqrt{r+z}$ and $\nu=\sqrt{r-z}$)
\begin{eqnarray}
& & \left[2\tilde E ({ \mu}^2+{ \nu}^2)
- {1\over 4}\mu^2\nu^2(\mu^2+\nu^2) + 4 \right] \, \Psi(\mu,\nu) \nonumber\\
&=& w^{-2} \left(\hat p_\mu^2 + \hat p_\nu^2\right) \, \Psi(\mu,\nu) \; ,
\label{H_eps}
\end{eqnarray}
with the radial operators ${{\hat { p}}_\mu}^2$ and
${{\hat { p}}_\nu}^2$ defined as
\[
\hat p_\mu^2
= -{1\over\mu}{\partial \over {\partial\mu}}
\left( \mu{\partial \over {\partial\mu}}\right) \; , \quad
\hat p_\nu^2
= -{1\over\nu}{\partial \over {\partial\nu}}
\left( \nu{\partial \over {\partial\nu}}\right) \; .
\]
Eq.\ \ref{H_eps} can be written in matrix form by using an appropriate
basis set.
The resulting generalized eigenvalue problem is solved numerically.
It has to be noted that the eigenvectors obtained, $|\psi_n\rangle$,
are orthonormal with respect to the scaled momentum operator, i.e.
\begin{equation}
\langle\psi_m|\hat p_\mu^2 + \hat p_\nu^2|\psi_n\rangle
= \langle m|n\rangle = \delta_{mn} \; ,
\end{equation}
with modified eigenvectors $|n\rangle$ defined by $|n\rangle\equiv
(\hat p_\mu^2+\hat p_\nu^2)^{1/2}|\psi_n\rangle$.
The diagonal matrix elements of an operator $\hat A$ are therefore obtained as
\begin{equation}
A_{nn} = \langle n|\hat A|n \rangle
= \langle\psi_n|\hat A (\hat p_\mu^2 + \hat p_\nu^2)|\psi_n \rangle \; .
\end{equation}
We are now prepared to compare the quantum spectra of the hydrogen atom in
a magnetic field with the semiclassical approximations in the chaotic and
regular regime of the classical phase space.
\subsubsection{Chaotic regime}
We have calculated 3181 eigenvalues $w_n<80$ of the scaling parameter and
the diagonal matrix elements of the two operators $\hat A=1/(2r{\bf p}^2)$
and $\hat B=r{\bf p}^2$ for the hydrogen atom in a magnetic field at
constant scaled energy $\tilde E=-0.1$.
The distributions of the matrix elements are presented in Fig.\ \ref{fig1}.
The matrix elements are distributed randomly around the mean values without
showing any regular pattern, as is typical of systems with chaotic
dynamics.
The quantum mechanical weighted density of states
\begin{equation}
\rho^{(A)}(w) = \sum_n \langle n|\hat A|n\rangle \delta(w-w_n)
\label{rho_A_qm_scaled:eq}
\end{equation}
can now be analyzed with the harmonic inversion technique to obtain the
scaled actions $s_p$ and the amplitudes $d^{(A)}_p$ (see Eq.\ \ref{semicl})
of the classical periodic orbits.
As can be seen from Eq.\ \ref{rho_A_scaled:eq} the periodic orbit amplitudes
\newpage
\phantom{}
\begin{figure}
\vspace{9.5cm}
\special{psfile=fig1.ps voffset=-25 hoffset=-25 vscale=46 hscale=46}
\caption{\label{fig1}
Values of the diagonal matrix elements $\langle n|\hat A|n\rangle$ for the
hydrogen atom in a magnetic field at scaled energy $\tilde E=-0.1$ in the
chaotic region of phase space as functions of the scaling parameter
$w=\gamma^{-1/3}$ ($\gamma\sim$ magnetic field strength):
(a) operator $\hat A=1/(2r{\bf p}^2)$; (b) $\hat B = r{\bf p}^2$.
}
\end{figure}
\begin{equation}
d^{(A)}_p = A_p d_p
\label{d_Ap:eq}
\end{equation}
are given as the product of the amplitudes, $d_p$ of Gutzwiller's original
trace formula, and the classical periodic orbit means $A_p$ in
Eq.\ \ref{Ap_s:eq}.
For the graphical presentation of the results it is therefore convenient
to divide the quantum amplitudes $d^{(A)}_p$ obtained by the
harmonic inversion of the spectra by the amplitudes, $d_p$ of
Gutzwiller's trace formula.
The periodic orbit quantities $A_p$ obtained in this way from the quantum
spectra at scaled energy $\tilde E=-0.1$ are presented in Fig.\ \ref{fig2}
for three different operators, viz.\ (a) the identity $\hat I$, and the
operators (b) $\hat A=1/(2r{\bf p}^2)$ and (c) $\hat B=r{\bf p}^2$.
The solid lines and crosses in Fig.\ \ref{fig2} mark the periodic orbit
means obtained by the harmonic inversion of the quantum spectra.
For comparison the dashed lines and squares present the periodic orbit
means of the observable obtained classically via Eq.\ \ref{Ap_s:eq}.
For the identity the classical periodic orbit averages
(squares in Fig.\ \ref{fig2}a) are exactly equal to one.
This is in excellent agreement with the harmonic inversion analysis of the
quantum mechanical density of states (crosses in Fig.\ \ref{fig2}a),
despite the two weakly separated periodic orbit contributions around
$s/2\pi\approx 1.1$.
For the two operators $\hat A=1/(2r{\bf p}^2)$ in Fig.\ \ref{fig2}b and
$\hat B=r{\bf p}^2$ in Fig.\ \ref{fig2}c the agreement between the
periodic orbit means obtained by harmonic inversion of the quantum
spectra and classically by Eq.\ \ref{Ap_s:eq} is of similar high accuracy as
\newpage
\phantom{}
\begin{figure}
\vspace{9.7cm}
\special{psfile=fig2.ps voffset=-25 hoffset=-25 vscale=46 hscale=46}
\caption{\label{fig2}
Periodic orbit means of observables (a) the identity, (b) $A=1/(2r{\bf p}^2)$,
and (c) $B = r{\bf p}^2$ for the hydrogen atom in a magnetic field at scaled
energy $\tilde E=-0.1$ as functions of the scaled action.
Solid lines and crosses: Results of the harmonic inversion of quantum spectra.
Dashed lines and squares: Periodic orbit means obtained by classical
calculations.
The agreement between the quantum and the classical calculations seems
to be excellent, except for the nearly degenerate recurrences at
$s/2\pi\approx 1.1$.
}
\end{figure}
\noindent
for the identity in Fig.\ \ref{fig2}a.
The results presented in Fig.\ \ref{fig2} demonstrate that harmonic inversion
of quantum spectra \cite{Mai97b} is indeed a powerful tool for the high
precision check of semiclassical theories.
Fig.\ \ref{fig2} provides an excellent numerical verification, by way of
example of the hydrogen atom in a magnetic field and the chosen set of
operators, of the validity of the semiclassical trace formula
(\ref{rho_A_scaled:eq}) for chaotic systems.
\subsubsection{Regular regime}
In the same way as described above we have checked the validity of the
extended Berry-Tabor formula (\ref{BT_A_scaled:eq}) for integrable systems.
As a physical system we again choose the hydrogen atom in a magnetic field,
but at low scaled energy $\tilde E=-0.5$, where the classical phase space
is regular.
We have calculated 5640 eigenvalues $w_n<160$ of the scaling parameter and
the diagonal matrix elements of the two operators $\hat A=1/(2r{\bf p}^2)$
and $\hat B=r{\bf p}^2$.
The weighted density of states (\ref{rho_A_qm_scaled:eq}) for the identity,
and the operators $\hat A=1/(2r{\bf p}^2)$ and $\hat B=r{\bf p}^2$ have been
analyzed in the same way as explained above.
The results obtained for the regular system at scaled energy $\tilde E=-0.5$
resemble those of Fig.\ \ref{fig2} for the chaotic system.
The difference is that the averages of the observables for the resonant tori
have been extracted from the quantum spectra by application of the
generalized Berry-Tabor formula (\ref{BT_A_scaled:eq}).
The quantum results perfectly agree with the classical averages which
illustrates the validity of the generalized Berry-Tabor formula.
\section{Non-trace type formulas}
\label{non-trace:sec}
The generalized semiclassical trace formulas (\ref{rho_A_scaled:eq})
and (\ref{BT_A_scaled:eq}) discussed in Sec.\ \ref{trace:sec} allow
the semiclassical calculation of the diagonal matrix elements of smooth
operators.
However, it would be desirable to know even more generalized
expressions for the calculation of {\em products} of matrix elements.
As mentioned in the introduction, such formulas are important, e.g., in
the semiclassical theory of matrix element fluctuations \cite{Eck95} or
for the construction of cross-correlated periodic orbit sums \cite{Mai99}.
To study matrix element fluctuations of an operator $\hat A$ the density
of states can be weighted with the variances
${\rm Var}_n A\equiv\langle n|\hat A^2|n\rangle-\langle n|\hat A|n\rangle^2$,
i.e.
\begin{eqnarray}
\rho^{({\rm Var}~A)}(E) &\equiv&
\sum_n \langle n|\hat A^2|n\rangle \delta(E-E_n) \nonumber \\
&-& \sum_n \langle n|\hat A|n\rangle^2 \delta(E-E_n) \; .
\label{rho_Var_A_qm:eq}
\end{eqnarray}
The first term in (\ref{rho_Var_A_qm:eq}) can be written as a semiclassical
trace formula (Eqs.\ \ref{rho_A_sc:eq} and \ref{BT_A:eq} for chaotic and
regular systems, respectively) with the observable $A$ replaced with its
square, $A^2$.
However, because of the squares of the matrix elements, the second
term in (\ref{rho_Var_A_qm:eq}) in general cannot be expressed in a
straightforward fashion with the help of the Green's operator $\hat G_E^+$
as a trace formula.
The trivial exception is when the operator $\hat A$ commutes with the
Hamiltonian, which means that $A$ is a constant of motion and thus its
variance vanishes.
Thus the derivation of a semiclassical approximation to the second term
in (\ref{rho_Var_A_qm:eq}) constitutes a nontrivial problem.
One solution can be obtained by application of periodic orbit sum rules
\cite{Ber85}.
Using smooth approximations of the $\delta$-functions, e.g.\ Gaussians
of width $\epsilon$,
\begin{equation}
\delta_\epsilon(E) = {1\over\sqrt{2\pi}\epsilon} e^{-E^2/2\epsilon^2} \; ,
\end{equation}
and the relation
\begin{equation}
\delta_\epsilon^2(E)
= {1\over 2\sqrt{\pi}\epsilon} \delta_{\epsilon/\sqrt{2}}(E)
\end{equation}
the second term in (\ref{rho_Var_A_qm:eq}) can formally be written as
the square of the density of states weighted with the diagonal matrix
elements \cite{Ber85,Eck95}, viz.
\begin{eqnarray}
& & \sum_n \langle n|\hat A|n\rangle^2
\delta_{\epsilon/\sqrt{2}} (E-E_n) \nonumber \\
&=& 2\sqrt{\pi}\epsilon\sum_n \langle n|\hat A|n\rangle^2
\delta_\epsilon^2 (E-E_n) \nonumber \\
&=& 2\sqrt{\pi}\epsilon\left[\sum_n \langle n|\hat A|n\rangle
\delta_\epsilon (E-E_n)\right]^2 \; .
\label{Berry_trick:eq}
\end{eqnarray}
The width $\epsilon$ in (\ref{Berry_trick:eq}) must be chosen sufficiently
small so that the smoothed $\delta$-functions do not overlap.
However, it should be noted that this condition cannot be fulfilled for
systems with degenerate states.
On the r.h.s.\ of Eq.\ \ref{Berry_trick:eq} the weighted density of states
can now be written as a trace formula and replaced with its semiclassical
approximations (\ref{rho_A_sc:eq}) and (\ref{BT_A:eq}) for chaotic and
regular systems, respectively.
Evaluation of the square of the periodic orbit sum on the r.h.s.\ of
(\ref{Berry_trick:eq}) then finally yields a double sum over all periodic
orbits of the classical system.
Although this result is formally correct, it is very inconvenient for
practical applications for the following reasons.
Firstly, the number of periodic orbits proliferates exponentially in
chaotic systems and the handling of the single periodic orbit sum is
already nontrivial.
The practical evaluation of the double sum would be even more cumbersome.
Secondly, the width $\epsilon$ in (\ref{Berry_trick:eq}) is a free
parameter.
Although the results should not depend on the width if $\epsilon$ is chosen
sufficiently small, the appropriate choice may render numerical calculations
extremely expensive.
Thirdly, and most importantly, the r.h.s.\ of Eq.\ \ref{Berry_trick:eq} does
not coincide with the ``simple'' trace formulas in those special cases,
where the operator $\hat A$ commutes with the Hamiltonian.
Even for the simplest operator, the identity $\hat A=\hat I$,
we end up with the nontrivial periodic orbit sum rule of Ref.\ \cite{Ber85}
instead of Gutzwiller's trace formula for the density of states.
Especially the third point indicates that the procedure described above
might not be the simplest way to construct a semiclassical approximation
to non-trace type formulas such as Eq.\ \ref{rho_Var_A_qm:eq}.
It is the main objective of this section to present a semiclassical
approximation to non-trace type weighted densities of states.
Our semiclassical expressions agree with the well established ``simple''
semiclassical trace formulas when the weighted density of states can be
written, for at least one of the operators commuting with the Hamiltonian,
as a quantum mechanical trace formula.
Starting from a more general equation than (\ref{rho_Var_A_qm:eq})
we study the density of states
\begin{equation}
\rho^{(A,B)}(E) \equiv \sum_n \langle n|\hat A|n\rangle
\langle n|\hat B|n\rangle \delta(E-E_n) \; ,
\label{rho_AB_qm:eq}
\end{equation}
weighted with the product of the diagonal matrix elements of two smooth
operators $\hat A$ and $\hat B$.
Eq.\ \ref{rho_AB_qm:eq} is the starting point to construct a quantum
mechanical cross-correlation function from a set of operators in Ref.\
\cite{Mai99}.
The variance of matrix elements (Eq.\ \ref{rho_Var_A_qm:eq}) is obtained
by setting $\hat B=\hat A$.
The weighted density of states (\ref{rho_AB_qm:eq}) can only be written
as a trace formula,
$\rho^{(A,B)}(E)=(-1/\pi)\,{\rm Im~tr}\,\{\hat A\hat G_E^+\hat B\}$
if either $\hat A$ or $\hat B$ commutes with the Hamiltonian.
As discussed in Sec.\ \ref{trace:sec} (see Eq.\ \ref{d_Ap:eq})
the semiclassical expressions for the weighted densities of states differ
from Gutzwiller's trace formula and the Berry-Tabor formula in the
following way.
The periodic orbit amplitudes are multiplied with the classical periodic
orbit (or torus) averages of the observable $A$.
We now assume that this ansatz is still valid for the non-trace type weighted
density of states (\ref{rho_AB_qm:eq}), i.e., its semiclassical analogue
has the same functional form as Gutzwiller's periodic orbit sum but with
periodic orbit amplitudes $d_p$ multiplied with the classical averages
$A_p$ and $B_p$ of both observables $A$ and $B$,
\begin{equation}
d^{(A,B)}_p = A_p B_p d_p \; ,
\label{d_ABp:eq}
\end{equation}
with $A_p$ and $B_p$ given by Eqs.\ \ref{Ap_t:eq} and \ref{Ap_reg:eq}
for chaotic and regular systems, respectively.
As can easily be seen, this ansatz has the property that the trace formulas
(\ref{rho_A_sc:eq}) and (\ref{BT_A:eq}) are recovered if one of the
operators is chosen to be the identity or one of the operators commutes
with the Hamiltonian.
However, the general validity of this ansatz is not at all obvious, and will
be checked numerically by the high resolution analysis of quantum spectra
in the following.
With the ansatz (\ref{d_ABp:eq}) for the periodic orbit amplitudes the
semiclassical analogue to the non-trace type formula (\ref{rho_AB_qm:eq})
reads
\begin{eqnarray}
\label{rho_AB_sc:eq}
& & \rho^{(A,B)}(E) = \rho^{(A,B)}_0(E) \\
&+& {1\over\pi\hbar}\, {\rm Re}\, \sum_p A_p B_p \sum_{r=1}^\infty
{T_p\over\sqrt{|\det(M_p^r-I)|}}\,e^{i[S_p(E)/\hbar-{\pi\over 2}\mu_p]r} \; ,
\nonumber
\end{eqnarray}
for systems with underlying chaotic classical dynamics, and
\begin{eqnarray}
\label{BT_AB:eq}
& & \rho^{(A,B)}(E) = \rho^{(A,B)}_0(E) \\
&+& {1\over\pi\hbar^{3/2}}\, {\rm Re}\, \sum_{\bf M} A_{\bf M} B_{\bf M}
{T_{\bf M}\over M_2^{3/2}|g_E''|^{1/2}} \,
e^{i[S_{\bf M}(E)/\hbar-{\pi\over 2}\eta_{\bf M}-{\pi\over 4}]}\, ,
\nonumber
\end{eqnarray}
for integrable systems.
Eqs.\ \ref{rho_AB_sc:eq} and \ref{BT_AB:eq} are the central propositions
of this paper, and generalize the semiclassical trace formulas
(\ref{rho_A_sc:eq}) and (\ref{BT_A:eq}) to the non-trace type weighted
density of states (\ref{rho_AB_qm:eq}).
The nontrivial statement is that weighting the quantum
mechanical density of states with the product of diagonal matrix elements
of smooth operators is equivalent, on the semiclassical level, to weighting
the periodic orbit contributions in the periodic orbit sum with the product
of the averages of the corresponding classical observables.
In analogy to the discussion of scaling properties in Sec.\ \ref{trace:sec},
Eqs.\ \ref{rho_AB_sc:eq} and \ref{BT_AB:eq} can be reformulated for scaling
systems, viz.
\begin{eqnarray}
\label{rho_AB_scaled:eq}
& & \rho^{(A,B)}(w)
= \rho^{(A,B)}_0(w) \\
&+& {1\over\pi\hbar}\, {\rm Re}\, \sum_p A_p B_p \sum_{r=1}^\infty
{s_p\over\sqrt{|\det(M_p^r-I)|}}\, e^{i[s_p w-{\pi\over 2}\mu_p]r}
\nonumber
\end{eqnarray}
for chaotic systems, and
\begin{eqnarray}
\label{BT_AB_scaled:eq}
& & \rho^{(A,B)}(w) = \rho^{(A,B)}_0(w) \\
&+& {1\over\pi\hbar^{3/2}}\, {\rm Re}\, \sum_{\bf M} A_{\bf M} B_{\bf M}
{s_{\bf M}\over M_2^{3/2}|g_E''|^{1/2}} \,
e^{i[s_{\bf M}w-{\pi\over 2}\eta_{\bf M}-{\pi\over 4}]}
\nonumber
\end{eqnarray}
for two-dimensional systems with regular dynamics.
For scaling systems the classical periodic orbit averages $A_p$ and $B_p$
in (\ref{rho_AB_scaled:eq}) must be calculated with respect to the
classical action instead of time as defined in Eq.\ \ref{Ap_s:eq}.
In the following we will provide convincing numerical evidence for the
validity of the semiclassical non-trace type formulas by the high precision
analysis (harmonic inversion) of quantum spectra of two different systems,
viz.\ the hydrogen atom in a magnetic field and the circle billiard.
A rigorous mathematical proof of the expressions given above is still
lacking and constitutes a challenge for the further development of
semiclassical theories.
\subsection{Hydrogen atom in a magnetic field}
To demonstrate the validity of the semiclassical non-trace type formulas,
Eqs.\ \ref{rho_AB_scaled:eq} and \ref{BT_AB_scaled:eq}, we use the same
system and set of operators as in Sec.\ \ref{trace:sec}, viz.\ the hydrogen
atom in a magnetic field at scaled energies $\tilde E=-0.1$ and
$\tilde E=-0.5$ in the chaotic and near-integrable regime, respectively,
and the operators $\hat A=1/(2r{\bf p}^2)$ and $\hat B=r{\bf p}^2$.
With the quantum mechanical eigenvalues and diagonal matrix elements at hand,
we construct the weighted densities of states (see Eq.\ \ref{rho_AB_qm:eq})
(a) $\rho^{(A,A)}(w)$, (b) $\rho^{(B,B)}(w)$, and (c) $\rho^{(A,B)}(w)$.
These spectra are analyzed with the harmonic inversion technique as described
in Sec.\ \ref{trace:sec}.
The analysis provides the scaled action $s_p$ of the periodic orbits and the
periodic orbit amplitudes $d^{(A,A)}_p$ ($d^{(B,B)}_p$ and $d^{(A,B)}_p$).
The amplitudes of the weighted densities of states are divided by the
amplitudes of the unweighted densities of states to obtain, according to
Eq.\ \ref{d_ABp:eq}, the products of the periodic orbit means $A_p^2$
($B_p^2$ and $A_pB_p$).
These values are presented as solid lines and crosses in Fig.\ \ref{fig5}
for the spectra in the chaotic regime at scaled energy $\tilde E=-0.1$ and
in Fig.\ \ref{fig6} for the spectra at scaled energy $\tilde E=-0.5$ in the
near-integrable regime.
For comparison, the squares mark the products of the periodic orbit means
obtained from the classical calculations.
As in Fig.\ \ref{fig2} for the high precision check of the
semiclassical trace formula (\ref{rho_A_scaled:eq}), the agreement between
the quantum and classical calculations is found to be very good, which
strongly supports the validity of the semiclassical non-trace type expressions.
Note that the somewhat larger deviations between the crosses and squares
for the nearly degenerate recurrencies at $s_p/2\pi\approx 1.1$ in Fig.\
\ref{fig5} have also been observed in Fig.\ \ref{fig2} for the semiclassical
trace formulas, i.e., the deviation does not indicate any failure of the
non-trace type formula (\ref{rho_AB_scaled:eq}).
\newpage
\phantom{}
\begin{figure}
\vspace{9.5cm}
\special{psfile=fig5.ps voffset=-25 hoffset=-25 vscale=46 hscale=46}
\caption{\label{fig5}
Products of periodic orbit means of the two observables $A=1/(2r{\bf p}^2)$
and $B = r{\bf p}^2$ for the hydrogen atom in a magnetic field at scaled
energy $\tilde E=-0.1$ as functions of the scaled action.
Solid lines and crosses: Results of the harmonic inversion of the non-trace
type weighted densities of states.
Dashed lines and squares: Results obtained by classical calculations.
As in Fig.\ \ref{fig2}, the agreement between the quantum and the classical
calculations seems to be excellent, except for the nearly degenerate
recurrences at $s/2\pi\approx 1.1$.
}
\end{figure}
\subsection{Circle billiard}
We now investigate the validity of the semiclassical non-trace type formula
(\ref{BT_AB_scaled:eq}) on a second system, viz.\ the integrable circle
billiard.
This system also serves as a model example in Ref.\ \cite{Mai99} to
construct a semiclassical cross-correlated periodic orbit sum for a given
set of smooth observables, and to calculate semiclassical spectra and
diagonal matrix elements by harmonic inversion of the cross-correlation
function.
As is well known, Schr\"odinger's equation for the circle billiard with
radius $R$ can be separated in polar coordinates $(r,\varphi)$, and the
wave functions can be expressed in terms of Bessel functions,
\begin{equation}
\psi_{nm}(r,\varphi)
= {\cal N}_{nm} J_{|m|}(k_{nm}r)e^{im\varphi} \; ,
\label{Bessel:eq}
\end{equation}
with the ${\cal N}_{nm}$ being normalization constants, $m$ the angular
momentum quantum number, and $k_{nm}=\sqrt{2ME_{nm}}/\hbar$ the quantized wave
numbers obtained as the $n$th zero of Bessel functions, $J_{|m|}(k_{nm}R)=0$.
In the following we choose radius $R=1$.
We calculated
\newpage
\phantom{}
\begin{figure}
\vspace{9.2cm}
\special{psfile=fig6.ps voffset=-35 hoffset=-25 vscale=46 hscale=46}
\caption{\label{fig6}
Same as Fig.\ 5 but at scaled energy $\tilde E=-0.5$.
}
\end{figure}
\noindent
31208 eigenvalues $k_{nm}<500$, and the diagonal matrix
elements of the operators $r$ and $r^2$.
The quantum spectra of (a) the unweighted density of states $\rho(k)$,
(the wave number $k$ is the scaling parameter, $w=k$ for billiard systems
\cite{Mai99}) and the density of states weighted with the matrix element
expressions
(b) $\langle\psi_{nm}|r^2|\psi_{nm}\rangle$,
(c) $\langle\psi_{nm}|r|\psi_{nm}\rangle^2$, and
(d) the variance $\langle{\rm Var}~r\rangle_{nm}\equiv
\langle\psi_{nm}|r^2|\psi_{nm}\rangle-\langle\psi_{nm}|r|\psi_{nm}\rangle^2$
have been analyzed with the harmonic inversion method.
The amplitudes obtained, divided by the amplitudes of the Berry-Tabor formula,
are presented as solid lines and crosses in Fig.\ \ref{fig7}, and the
corresponding classical averages are drawn as squares for comparison.
As can be seen, the agreement is perfect, not only for the identity and the
periodic orbit means of the observable $r^2$ in Fig.\ \ref{fig7}a and
\ref{fig7}b, verifying the Berry-Tabor formula and its extension
(\ref{BT_A_scaled:eq}), but also for the squares of the periodic orbit means
of $r$ and the variance of this observable in Fig.\ \ref{fig7}c and
\ref{fig7}d, where the agreement demonstrates the validity of the non-trace
type equation (\ref{BT_AB_scaled:eq}) for the circle billiard with
$\hat A=\hat B=r$.
The squares in Fig.\ \ref{fig7}d mark the classical variances of
the observable $r$ on the various resonant tori.
Our conjecture therefore provides a basic formula for semiclassical matrix
element fluctuations, since it directly relates the quantum variances
${\rm Var}_n A\equiv\langle n|\hat A^2|n\rangle-\langle n|\hat A|n\rangle^2$
of a smooth operator $\hat A$ to the classical variances
${\rm Var}_p A\equiv \langle A^2\rangle_p-\langle A\rangle_p^2$
of the observable $A$ taken along the periodic orbits or resonant tori.
The perfect agreement between the quantum and classical results for the
circle billiard in Fig.\ \ref{fig7} compared to
\newpage
\phantom{}
\begin{figure}
\vspace{10.5cm}
\special{psfile=fig7.ps voffset=-25 hoffset=-25 vscale=46 hscale=46}
\caption{\label{fig7}
Classical averages on rational tori of (a) the identity,
(b) $\langle r^2\rangle_p$, (c) $\langle r\rangle_p^2$, and (d) the variances
${\rm Var}_p r\equiv\langle r^2\rangle_p-\langle r\rangle_p^2$
for the circle billiard with radius $R=1$ as functions of the scaled action.
Solid lines and crosses: Results of the harmonic inversion of quantum spectra.
Dashed lines and squares: Periodic orbit means obtained by classical
calculations.
}
\end{figure}
\noindent
the very good but not
absolutely perfect results for the hydrogen atom in a magnetic field
in Figs.\ \ref{fig5} and \ref{fig6} may be explained by the different
number of quantum states used in the harmonic inversion analysis.
For the circle billiard we have calculated more than 30000 states, which
is by about a factor of 10 (5.5) times more quantum states than for
the hydrogen atom in a magnetic field at scaled energy $\tilde E=-0.1$
($\tilde E=-0.5$).
\section{Conclusion and outlook}
\label{conclusion}
We have extended semiclassical trace formulas for the density of states
of regular and chaotic systems, or the density of states weighted with
the diagonal matrix elements of smooth operators, to the more general
class of {\em non-trace} type equations, where the density of states
is weighted with the diagonal matrix elements of two operators $\hat A$
and $\hat B$, i.e., $\rho^{(A,B)}(E)
=\sum_n\langle n|\hat A|n\rangle\langle n|\hat B|n\rangle\delta(E-E_n)$.
By the high resolution analysis (harmonic inversion) of the
quantum spectra of two different systems, viz.\ the hydrogen atom in a
magnetic field and the circle billiard, we have given numerical evidence
that weighting the quantum mechanical density of states with the product
of the diagonal matrix elements
$\langle n|\hat A|n\rangle\langle n|\hat B|n\rangle$
is equivalent, on the semiclassical level, to weighting the periodic orbit
contributions in the periodic orbit sum with the product of the averages
of the corresponding classical observables,
$\langle A\rangle_p\langle B\rangle_p$, where the means are taken along
the periodic orbits or resonant tori for chaotic and regular systems,
respectively.
However, a rigorous mathematical derivation of semiclassical non-trace
type formulas appears nontrivial, and would be a challenging task for the
further development of semiclassical theories.
There are several useful and important applications of semiclassical
non-trace type formulas.
For example, it enables the semiclassical approach to matrix element
fluctuations. The variances
${\rm Var}_n A\equiv\langle n|\hat A^2|n\rangle-\langle n|\hat A|n\rangle^2$
of the diagonal matrix elements of a smooth operator $\hat A$ are expressed
in terms of the variances
${\rm Var}_p A\equiv\langle A^2\rangle_p-\langle A\rangle_p^2$
of the classical observable $A$ taken along the periodic orbits or
resonant tori.
Non-trace type formulas also provide the semiclassical approximation to
cross-correlated weighted density of states,
$\rho_{\alpha\alpha'}(E)=\sum_n\langle n|\hat A_\alpha|n\rangle
\langle n|\hat A_{\alpha'}|n\rangle\delta(E-E_n)$
with a set of smooth operators $\hat A_\alpha$, $\alpha=1,\dots,D$.
The additional classical information obtained from the set of classical
observables can be used to significantly improve the convergence properties
of semiclassical quantization methods \cite{Mai99}.
In this paper we have investigated non-trace type expressions for products
of two diagonal matrix elements.
These products have been chosen because of the important applications
to semiclassical matrix element fluctuations, i.e., the calculation of
variances of matrix elements and to the semiclassical quantization method
in Ref.\ \cite{Mai99}.
However, our conjecture is not restricted to products of two matrix elements.
For example, it appears straightforward to generalize Eqs.\ \ref{rho_AB_sc:eq}
and \ref{BT_AB:eq} to products of more than two matrix elements and classical
periodic orbit means.
The most general case of non-trace type equations would be the analysis of
functions $f(A_{nn},B_{nn},C_{nn},\dots)$ of one or more diagonal matrix
elements, i.e.,
$\rho^{(f)}(E)=\sum_nf(A_{nn},B_{nn},C_{nn},\dots)\delta(E-E_n)$,
which should be obtained semiclassically by multiplying the
periodic orbit amplitudes of Gutzwiller's trace formula or the Berry-Tabor
formula with the function
$f(\langle A\rangle_p,\langle B\rangle_p,\langle C\rangle_p,\dots)$
of the periodic orbit means of the corresponding classical observables.
Certainly the operators and the function $f$ must be smooth.
Clearly, further investigations will be necessary to verify that conjecture
and to specify the smoothness conditions on operators and functions.
In conclusion, the analysis of non-trace type equations will provide a
valuable instrument for extending the relation between quantum mechanical
matrix elements on the one side and the periodic orbit means of classical
observables on the other.
\acknowledgements
We acknowledge stimulating discussions with J.\ Keating.
This work was supported in part by the Son\-der\-for\-schungs\-be\-reich
No.\ 237 of the Deutsche For\-schungs\-ge\-mein\-schaft.
J.M.\ is grateful to Deutsche For\-schungs\-ge\-mein\-schaft for a
Habilitandenstipendium (Grant No.\ Ma 1639/3).\\[-3ex]
|
\section{Introduction}
The determination of the value of the Hubble constant, $H_0$,
is one of the classical tasks of {\it observational}
cosmology. In the framework of the expanding space paradigm it
provides a measure of the distance scale in FRW universes and
its reciprocal gives the time scale.
This problem has been approached in various ways.
A review on the recent determinations of the value of $H_0$
shows that most methods provide values at $H_0\sim55\dots75$
(for brevity we omit the units; all $H_0$ values are in
$\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$): Virgo cluster yields
$55\pm7$ and clusters from Hubble diagram with relative
distances to Virgo $57\pm7$ (Federspiel et al.
\cite{Federspiel98}), type Ia supernovae give
$60\pm10$ (Branch \cite{Branch98}) or $65\pm7$ (Riess et
al. \cite{Riess98}), Tully-Fisher relation in I-band yields
$69\pm5$ (Giovanelli et al. \cite{Giovanelli97}) and
$55\pm7$ in B-band (Theureau et al. \cite{Theureau97b}, value
and errors combined from the diameter and magnitude
relations), red giant branch tip gives $60\pm11$
(Salaris \& Cassisi \cite{Salaris98}), gravitational lens
time delays $64\pm13$ (Kundi\'c et al. \cite{Kundic97})
and the `sosies' galaxy method $60\pm10$ (Paturel
et al. \cite{Paturel98}). Sunyaev-Zeldovich effect
has given lower values, $49\pm29$ by Cooray
(\cite{Cooray98}), $47^{+23}_{-15}$ by Hughes \&
Birkinshaw (\cite{Hughes98}), but the uncertainties in
these results are large due to various systematical
effects (Cen, \cite{Cen98}). Surface brightness
fluctuation studies provide a higher value of
$87\pm11$ (Jensen et al., \cite{Jensen99}), but
most methods seem to fit in the range 55 - 75
stated above. An important comparison to these local
values
may be found after the cosmic microwave background
anisotropy probes (MAP and Planck) and galaxy
redshift surveys (2dF and SDSS) offer us a
multitude of high resolution data (Eisenstein
et al., \cite{Eisenstein98}). Note that most of
the errors cited here as well as given in the
present paper are $1\sigma$ errors.
The present line of research has its roots in the work of
Bottinelli et al. (\cite{Bottinelli86}), where $H_0$ was
determined using spiral galaxies in the {\it field}. They
used the direct Tully-Fisher relation (Tully \&
Fisher \cite{Tully77}):
\begin{equation}
\label{E1}
\\ M\propto\log V_\mathrm{max},
\end{equation}
where $M$ is the absolute magnitude in a given band
and $\log V_\mathrm{max}$ is the maximum rotational
velocity measured from the hydrogen 21 cm line width of each galaxy.
Gouguenheim (\cite{Gouguenheim69})
was the first to suggest that such a relation might exist
as a distance indicator.
Bottinelli et al. (\cite{Bottinelli86}) paid particular attention
to the elimination of the so-called {\it Malmquist bias}. In
general terms, the determination of $H_0$ is subject to the
Malmquist bias of the $2^\mathrm{nd}$ kind:
the inferred value of $H_0$ depends on the distribution
of the derived distances $r$ for each true distance $r'$
(Teerikorpi \cite{Teerikorpi97}).
Consider the expectation value of the derived distance $r$
at a given true distance $r'$:
\begin{equation}
\label{E2}
\\ E(r\vert r')=\int\limits_0^{\infty}\!\mathrm{d}r\,r\,P(r\vert r').
\end{equation}
The integral is done over {\it derived} distances $r$.
For example, consider a strict magnitude
limit: for each true distance the derived distances are exposed to an
upper cut-off. Hence the expectation value for the derived
distance $r$ at $r'$ is too small and thus
$H_0$ will be overestimated.
Observationally, the direct Tully-Fisher relation takes the form:
\begin{equation}
\label{E3}
\\ X = \mathrm{slope}\times p + \mathrm{cst},
\end{equation}
where we have adopted a shorthand
$p$ for $\log V_\mathrm{max}$ and $X$ denotes either the
absolute magnitude $M$ or $\log D$, where $D$ labels the
absolute linear size of a galaxy in kpc.
In the {\it direct} approach the slope is determined from
the linear regression of $X$ against $p$. The resulting
direct Tully-Fisher relation can be expressed as
\begin{equation}
\label{E4}
\\ E(X\vert p) = ap+b.
\end{equation}
Consider now the {\it observed} average of $X$ at each $p$,
$\langle X\rangle_p$, as a function of the true distance.
The limit in $x$ (the observational counterpart of $X$)
cuts off progressively more and more of the distribution
function of $X$ for a constant $p$.
Assuming $X=\log D$ one finds:
\begin{equation}
\label{E5}
\\ \langle X\rangle_p \ge E(X\vert p),
\end{equation}
The inequality gives a practical measure of the Malmquist
bias depending primarily on $p$, $r'$, $\sigma_X$
and $x_\mathrm{lim}$.
The equality holds only when the $x$-limit cuts the
luminosity function $\Phi(X)$ insignificantly.
That the direct relation is {\it inevitably} biased by its
nature forces one either to look for an unbiased subsample
or to find an appropriate correction for the bias. The
former was the strategy chosen by Bottinelli et al.
(\cite{Bottinelli86}) where the method of normalized distances
was introduced. This is the method chosen also by the KLUN project.
KLUN ({\sl Kinematics of the Loal Universe}) is based on a
large sample, which consists of
5171 galaxies of Hubble types T=1-8 distributed on the whole
celestial sphere (cf. e.g. Paturel \cite{Paturel94},
Theureau et al. \cite{Theureau97b}).
Sandage (\cite{Sandage94a}, \cite{Sandage94b})
has also studied the latter approach. By recognizing that the
Malmquist bias depends not only on the imposed $x$-limit
but also on the rotational velocities and distances, he
introduced the triple-entry correction method,
which has consistently predicted values of $H_0$ supporting
the long cosmological distance scale. As a practical
example of this approach to the Malmquist bias cf.
e.g. Federspiel et al. (\cite{Federspiel94}).
Bottinelli et al. (\cite{Bottinelli86}) found
$H_0=72\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$
using the method of normalized distances, i.e. using a sample
cleaned of galaxies suffering from the Malmquist bias.
This value was based on the de Vaucouleurs calibrator distances.
If, instead, the Sandage-Tammann calibrator distances were used
Bottinelli et al. (\cite{Bottinelli86}) found
$H_0=63\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$
(or $H_0=56\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$
if using the old ST calibration).
One appreciates the debilitating effect of the Malmquist
bias by noting that when it is ignored
the de Vaucouleurs calibration yields much larger values:
$H_0\sim100\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$.
Theureau et al. (\cite{Theureau97b})
by following the guidelines set out by Bottinelli et
al. (\cite{Bottinelli86}) determined the value of $H_0$
using the KLUN sample.
$H_0$ was determined not only using magnitudes
but also diameters because the
KLUN sample is constructed to be complete in angular
diameters rather than magnitudes (completeness limit
is estimated to be $D_{25}=1\farcm6$).
Left with 400 unbiased galaxies (about ten times more than
Bottinelli et al. (\cite{Bottinelli86}) were able to use) reaching
up to $2000{-}3000\mathrm{\, km\, s^{-1}}$
they found using the most recent calibration based on HST
observations of extragalactic cepheids
\begin{itemize}
\item $H_0=53.4\pm 5.0\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$
from the magnitude relation, and
\item $H_0=56.7\pm 4.9\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$
from the diameter relation.
\end{itemize}
They also discussed in their Sect. 4.2 how these results
change if the older calibrations were used. For example,
the de Vaucouleurs calibration would increase these values
by 11 \%. We expect that a similar effect would be
observed also in the present case.
In the present paper we ask whether the results of
Theureau et al. (\cite{Theureau97b}) could be confirmed
by implementing the {\it inverse} Tully-Fisher relation:
\begin{equation}
\label{E6}
\\ p=a'X+b',
\end{equation}
This problem has special importance because of the ``unbiased''
nature that has often been ascribed to the inverse
Tully-Fisher relation as a distance
indicator and because of the large number of galaxies available contrary to
the direct approach where one is constrained to the
so called unbiased plateau (cf. Bottinelli et al. \cite{Bottinelli86};
Theureau et al. \cite{Theureau97b}). The fact that the inverse relation
has it own particular biases has received increasing attention
during the years (Fouqu\'e et al.
\cite{Fouque90}, Teerikorpi \cite{Teerikorpi90}, Willick
\cite{Willick91}, Teerikorpi \cite{Teerikorpi93},
Ekholm \& Teerikorpi \cite{Ekholm94}, Freudling et al.
\cite{Freudling95}, Ekholm \& Teerikorpi \cite{Ekholm97},
Teerikorpi et al. \cite{Teerikorpi99} and, of course, the
present paper).
\section{Outlining the approach}
As noted in the introduction the KLUN project approaches the
problem of the determination of the value of $H_0$ using
field galaxies with photometric distances. Such an approach
reduces to three steps
\begin{enumerate}
\item construction of a relative kinematical distance scale,
\item construction of a relative redshift-independent
distance scale, and
\item establishment of an absolute calibration.
\end{enumerate}
Below we comment on the first two steps. In particular we
further develop the concept of a relevant inverse slope
which may differ from the theoretical slope, but is still
the slope to be used. The third step is addressed in Sect.~6.
It is hoped that this review clarifies the methodological basis
of the KLUN project and also makes the notation used more familiar.
\subsection{The kinematical distance scale}
The first step takes its simplest form by assuming
the strictly linear Hubble law:
\begin{equation}
\label{E7}
\\ R_\mathrm{kin} = V_\mathrm{o}/H_0'
\end{equation}
where $V_\mathrm{o}$ is the radial velocity inferred from
the observed redshifts and $H_0'$ is some input value
for the Hubble constant. Because $V_\mathrm{o}$ reflects
the true kinematical distance $R_\mathrm{kin}^*$ via the
true Hubble constant $H_0^*$
\begin{equation}
\label{E8}
\\ R_\mathrm{kin}^* = V_\mathrm{o}/H_0^*,
\end{equation}
one recognizes that Eq.~\ref{E7} sets up a {\it relative}
distance scale:
\begin{equation}
\label{E9}
\\ d_\mathrm{kin} = \frac {R_\mathrm{kin}}{R_\mathrm{kin}^*}
= \frac {H_0^*}{H_0'}.
\end{equation}
In other words, $\log d_\mathrm{kin}$ is known next to
a constant.
In a more realistic case one ought to consider also the
{\it peculiar} velocity field. In KLUN one
assumes that peculiar velocities are governed
mainly by the Virgo supercluster.
In KLUN the kinematical distances are inferred
from $V_\mathrm{o}$'s by implementing the spherically
symmetric model of Peebles (\cite{Peebles76}) valid
in the linear regime. In the adopted form of this
model (for the equations to be solved cf. e.g.
Bottinelli et al. \cite{Bottinelli86}, Ekholm \cite{Ekholm96})
the centre of the peculiar velocity field is marked by the pair
of giant ellipticals M86/87 positioned at some unknown true distance
$R^*$ which is used to normalize the kinematical distance
scale: the centre is at a distance $d_\mathrm{kin}=1$.
The required cosmological velocities $V_\mathrm{cor}$
(observed velocities corrected for peculiar motions)
are calculated as
\begin{equation}
\label{E10}
\\ V_\mathrm{cor} = C_1\times d_\mathrm{kin},
\end{equation}
where the constant $C_1$ defines the linear recession
velocity of the centre of the system assumed to be
at rest with respect to the quiescent Hubble flow:
\begin{equation}
\label{E11}
\\ C_1 = V_\mathrm{o}(\mathrm{Vir})+
V_\mathrm{inf}^\mathrm{LG}.
\end{equation}
$V_\mathrm{o}(\mathrm{Vir})$ is the presumed velocity of the
centre and $V_\mathrm{inf}^\mathrm{LG}$ is the presumed
infall velocity of the Local Group into the centre of
the system.
\subsection{The redshift-independent distances}
The direct Tully-Fisher relation is quite
sensitivite to the sampling of the luminosity function.
On the other hand, when implementing the inverse Tully-Fisher
relation (Eq.~\ref{E6}) under ideal conditions
it does not matter how we sample $X$
(Schechter \cite{Schechter80}) in order to obtain an
unbiased estimate for the inverse parameters and,
furthermore, the expectation value
$E(r\vert r')$ is also unbiased (Teerikorpi \cite{Teerikorpi84}).
{\it However, we should sample all
$\log V_\mathrm{max}$ for each constant {\it true} $X$ in the sample}.
This theoretical prerequisition is often tacitly assumed
in practice. For more formal treatments on the inverse relation
cf. Teerikorpi(\cite{Teerikorpi84}, \cite{Teerikorpi90},
\cite{Teerikorpi97}) and e.g.
Hendry \& Simmons (\cite{Hendry94}) or Rauzy \& Triay (\cite{Rauzy96}).
In the inverse approach the distance indicator is
\begin{equation}
\label{E12}
\\ X = A'\langle p\rangle_X+\mathrm{cst.},
\end{equation}
where $A'=1/a'$ following the notation adopted by Ekholm
\& Teerikorpi (\cite{Ekholm97}; hereafter ET97).
The inverse regression slope $a'$ is expected to fulfill
\begin{equation}
\label{E13}
\\ \langle p\rangle_X \equiv E(p\vert X)=a'X+\mathrm{cst}.
\end{equation}
$\langle p\rangle_X$ is the observed average $p$ for
a given $X$. Eq.~\ref{E13} tells that in order to find
the correct $a'$ one must sample the distribution function
$\phi_X(p)$ in such a way that
$\langle p\rangle_X=(p_0)_X$, where $(p_0)_X$ is the
central value of the underlying distribution function.
$\phi_X(p)$ is presumed to be {\it symmetric}
about $(p_0)_X$ for all $X$.
ET97 demonstrated
how under these ideal conditions the derived $\log H_0$
as a function of the kinematical distance should run
horizontally as the adopted slope approaches the ideal,
theoretical slope.
In practice the parameters involved are subject to uncertainties,
in which case one should use instead of the unknown theoretical slope
a slope which we call the {\it relevant} inverse slope.
We would like to clarify in accurate terms the meaning of this slope
which differs from the theoretical slope
and which has been more heuristically discussed by
Teerikorpi et al. (\cite{Teerikorpi99}).
The difference between the theoretical and the relevant slope can be
expressed in the following formal way. Define the observed parameters
as
\begin{equation}
\label{E14}
\\ X_\mathrm{o}=X+\epsilon_x+\epsilon_\mathrm{kin},
\end{equation}
\begin{equation}
\label{E15}
\\ p_\mathrm{o}=p+\epsilon_p,
\end{equation}
where $X$ is inferred from $x$ with a measurement error $\epsilon_x$
and the kinematical distance $d_\mathrm{kin}$ has an error $\epsilon_\mathrm{kin}$ due to
uncertainties in the kinematical distance scale.
$\epsilon_p$ is the observational error on $p$.
The theoretical slope $a'_\mathrm{t}$ is\footnote
{We make use of the
formal definition of the slope of the linear regression of $y$
against $x$ with
\begin{equation}
\label{E16}
\\ \mathrm{Cov}(x,y) =
\frac{\sum(x-\langle x\rangle)(y-\langle y\rangle)}{(N-1)}.
\end{equation}
}
\begin{equation}
\label{E17}
\\ a'_\mathrm{t}=
\frac {\mathrm{Cov}(X,p)}{\mathrm{Cov}(X,X)},
\end{equation}
while the observed slope is
\begin{equation}
\label{E18}
\\ a'_\mathrm{o} =
\frac {\mathrm{Cov}(X_\mathrm{o},p_\mathrm{o})}{\mathrm{Cov}
(X_\mathrm{o},X_\mathrm{o})}
\sim
\frac {\mathrm{Cov}(X,p)+\mathrm{Cov}(\epsilon_x+\epsilon_\mathrm{kin},
\epsilon_p)}
{\mathrm{Cov}(X,X)+\sigma^2_x+\sigma^2_\mathrm{kin}}
\end{equation}
We call the slope $a'_\mathrm{o}$ relevant if it verifies for
{\it all} $X_\mathrm{o}$
(Eq.~\ref{E13})
\begin{equation}
\label{E19}
\\ \langle p_\mathrm{o}\rangle_{X_\mathrm{o}}=E(p\vert X_\mathrm{o})=
a'_\mathrm{o}X_\mathrm{o}+\mathrm{cst}.
\end{equation}
This definition means that
the average observed value of $p_\mathrm{o}$ at
each fixed value of $X_\mathrm{o}$ (derived from observations and
the kinematical distance scale) is correctly predicted by
Eq.~\ref{E19}.
Note also that in the case of diameter relation, $\epsilon_x$,
$\epsilon_\mathrm{kin}$ and $\epsilon_p$ are only weakly correlated.
Thus the difference between the relevant slope and the theoretical
slope is dominated by $\sigma_x^2+\sigma_\mathrm{kin}^2$. In the
special case where the galaxies are in one cluster (i.e. at the same
true distance), the dispersion $\sigma_\mathrm{kin}$
vanishes. In order to make the relevant slope more
tangible we demonstrate in Appendix A how
it indeed is the one to be
used for the determination of $H_0$.
Finally, also selection in $p$ and type effect may affect the
derived slope making it even shallower.
Theureau et al. (\cite{Theureau97a})
showed that a type effect exists seen as
degenerate values of $p$ for each constant linear diameter $X$.
Early Hubble types rotate faster than late types. In addition,
based on an observational program of 2700 galaxies with the
Nan\c{c}ay radiotelescope, Theureau (\cite{Theureau98})
warned that the detection
rate in HI varies continuosly from early to late types and that on
average $\sim 10\%$ of the objects remain unsuccessfully observed.
Influence of such a selection, which concerns principally the extreme
values of the distribution function $\phi(p)$,
was discussed analytically by
Teerikorpi et al. (\cite{Teerikorpi99}).
\section{A straightforward derivation of $\log H_0$}
\subsection{The sample}
KLUN sample is -- according to
Theureau et al. (\cite{Theureau97b}) -- complete up to
$B_\mathrm{T}^\mathrm{c}=13\fm25$, where $B_\mathrm{T}^\mathrm{c}$
is the corrected total B-band magnitude and down to
$\log D_{25}^\mathrm{c}=1.2$, where $D_{25}^\mathrm{c}$ is
the corrected angular B-band diameter.
The KLUN sample was subjected
to exclusion of low-latitude ($\vert b\vert\ge15\degr$)
and face-on ($\log R_{25}\ge0.07$) galaxies.
The centre of the
spherically symmetric peculiar velocity field was positioned
at $l=284\degr$ and $b=74\degr$. The constant $C_1$
needed in Eq.~\ref{E10} for cosmological velocities
was chosen to be $1200\mathrm{\, km\, s^{-1}}$ with
$V_\mathrm{o}(\mathrm{Vir})=980\mathrm{\, km\, s^{-1}}$
and $V_\mathrm{inf}^\mathrm{LG}=220\mathrm{\, km\, s^{-1}}$
(cf. Eq.~\ref{E11}). After the exclusion of triple-valued
solutions to the Peebles' model and
when the photometric completeness limits cited were imposed on
the remaining sample one was left with 1713 galaxies for
the magnitude sample and with 2822 galaxies for the diameter
sample.
\subsection{The inverse slopes and calibration of zero-points}
Theureau et al. (\cite{Theureau97a})
derived a common inverse diameter slope $a'\approx0.50$
and inverse magnitude slope $a'\approx-0.10$ for
all Hubble types considered i.e. T=1-8.
These slopes were also shown to obey a
a simple mass-luminosity model
(cf. Theureau et al. \cite{Theureau97a}).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f1}}
\caption{
The slope $a'=0.50$ forced to the calibrator
sample with Cepheid distances yielding $b'_{\mathrm{cal}}=1.450$,
when no type corrections were made.
}
\label{F1}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f2}}
\caption{
The slope $a'=-0.10$ forced to the calibrator
sample with Cepheid distances yielding $b'_\mathrm{cal}=0.117$,
when no type corrections were made.
}
\label{F2}
\end{figure}
With these estimates for the inverse
slope the relation can be calibrated.
At this point of derivation we ignore the
effects of type-dependence and possible selection in
$\log V_\mathrm{max}$.
The calibration was
done by forcing the slope to the calibrator sample
of 15 field galaxies with cepheid distances,
mostly from the HST programs (Theureau et al.
\cite{Theureau97b}, cf. their Table~1.).
The absolute zero-point is given by
\begin{equation}
\label{E20}
\\ b'_\mathrm{cal}=\frac
{\sum {(\log V_\mathrm{max}-a'X)}}{N_\mathrm{cal}},
\end{equation}
where the adopted inverse slope $a'=0.50$ yields $b'_\mathrm{cal}=1.450$
and $a'=-0.10$ $b'_\mathrm{cal}=0.117$. In Fig.~\ref{F1} we show
the calibration for the diameter relation and in Fig.~\ref{F2}
for the magnitude relation.
\subsection{$H_0$ without type corrections}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,123][460,507]{8156.f3}}
\caption{Panel (a): The $\log H_0$ vs. $V_\mathrm{cor}$ diagram for the
calibrated inverse Tully-Fisher relation $\log V_\mathrm{max}=
0.50\log D+1.450$. The horizontal solid line corresponds
to the average value $\langle\log H_0\rangle=1.92$. Panel (b):
the average values $\langle\log H_0\rangle$ (circles) are shown as
well as the average of the whole sample. The averages were calculated
for velocity bins of size $1000\mathrm{\ km\,s^{-1}}$. Total number
of points used was $N=2822$.}
\label{F3}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,123][460,507]{8156.f4}}
\caption{The sample imposed to the strict magnitude limit
$B_\mathrm{T}^\mathrm{c}=13\fm25$ (N=1713). The forced
solution yields
$\langle\log H_0\rangle=1.857$ or
$H_0=71.9\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$.
}
\label{F4}
\end{figure}
ET97 discussed in some detail problems which hamper the determination
of the Hubble constant $H_0$ when one applies the inverse Tully-Fisher
relation. They concluded that
once the relevant inverse slope is found, the
average $\langle\log H_0\rangle$ shows no tendencies as a function
of the distance. Or, in terms of the method of normalized distances
of Bottinelli et. al. (\cite{Bottinelli86}),
the unbiased plateau extends to all
distances. ET97 also noted how one might simultaneously fine-tune
the inverse slope and get an unbiased estimate for $\log H_0$.
The resulting $\log H_0$ vs. kinematical distance diagrams
for the inverse diameter relation is given in
Fig.~\ref{F3} and for the magnitude relation in Fig.~\ref{F4}.
Application of the parameters given in the previous section yield
$\langle\log H_0\rangle=1.92$ correponding to
$H_0=83.2\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$ for the diameter sample
and
$\langle\log H_0\rangle=1.857$ or
$H_0=71.9\mathrm{\, km\, s^{-1}\, Mpc^{-1}}$
for the magnitude sample.
These averages are shown as horizontal, solid
straight lines. In panels (a) individual points are plotted and
in panels (b) the averages for bins of
$1000\mathrm{\, km\, s^{-1}}$ are given as circles.
Consider first the diameter relation. One clearly
sees how the average follows a
horizontal line up to $9000\mathrm{\ km\,s^{-1}}$.
At larger distances, the observed behaviour of
$\langle H_0\rangle$ probably
reflects some selection in $\log V_\mathrm{max}$ in the sense that
there is an upper cut-off value for $\log V_\mathrm{max}$.
Note also the mild downward tendency between $1000\mathrm{\, km\, s^{-1}}$
and $5000\mathrm{\, km\, s^{-1}}$.
Comparison of Fig.~\ref{F4} with Fig.~\ref{F3}
shows how $\langle\log H_0\rangle$ from magnitudes
and diameters follow each other quite well as expected
(ignoring, of course, the vertical shift in the averages).
Note how the growing tendency of
$\langle\log H_0\rangle$ beyond $9000\mathrm{\, km\, s^{-1}}$
is absent in the magnitude sample
because of the limiting magnitude:
the sample is less deep. This suggests that
the possible selection bias in $\log V_\mathrm{max}$
does not affect the magnitude sample.
One might, by the face-value,
be content with the slopes adopted as well as with
the derived value of $H_0$. The observed behaviour is what ET97
argued to be the prerequisite for an unbiased estimate for
the Hubble constant: non-horizontal trends disappear.
It is -- however -- rather disturbing to note that the values of
$H_0$ obtained via this straightforward application of the inverse
relation are significantly {\it larger} than those reported
by Theureau et al. (\cite{Theureau97b}). The inverse diameter
relation predicts some
50 percent larger value and the magnitude relation some 30
percent larger value than the corresponding direct relations.
In what follows, we try to understand this discrepancy.
\section{Is there selection in $\log V_\mathrm{max}$?}
The first explanation coming to mind is that the apparently
wellbehaving slope $a'= 0.5$ ($a'=-0.1$) is incorrect because
of some selection effect and is thus {\it not} relevant in the
sense discussed in Sect.~2.2 and in Appendix A.
The relevant slope brings about
an unbiased estimate for the Hubble parameter (or the Hubble
constant if one possesses an ideal calibrator sample) {\it if}
the distribution function of $\log V_\mathrm{max}$, $\phi(p)_X$,
is completely and correctly sampled for each $X$.
Fig.~\ref{F3} showed some preliminary indications that
this may not be the case as regards the diameter sample.
Teerikorpi (\cite{Teerikorpi99}) discussed the effect and
significance of a strict upper and/or lower cut-off on
$\phi(p)_X$. For example, an upper cut-off in $\phi(p)_X$
should yield a too large value of $H_0$ and, furthermore,
a too shallow slope. Their analytical calculations given
the gaussianity of $\phi(p)_X$ show that this kind of selection
effect has only a minuscule affect unless the cut-offs are
considerable. Because the selection does not seem to be
significant, we do not expect much improvement in $H_0$.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f5}}
\caption{The inverse Tully-Fisher diagram for the sample used
in the analysis. The solid line refers to a linear regression
of $a'=0.576$ and $b'=1.256$. The dashed lines give the forced
solutions with $a'=0.50$ for Hubble types 1 with $b'=1.448$
and 8 with $b'=1.209$
The dotted lines at
$\log V_\mathrm{max}=2.55$ and $\log V_\mathrm{max}=1.675$ are intended
to guide the eye. At least the upper cut-off is quite conspicuous.}
\label{F5}
\end{figure}
There is, however, another effect which may alter the slope.
As mentioned in Sect.~2.2 the type-dependence of the zero-point
should be taken into account. Because the selection function
may depend on the morphological type it also affects the type
corrections. This is clearly seen when one considers how the
type corrections are actually calculated. As in Theureau et al.
(\cite{Theureau97b}) galaxies are shifted to a common Hubble type 6
by applying a correction term $\Delta b' = b'(T)-b'(6)$ to individual
$\log V_\mathrm{max}$ values, where
\begin{equation}
\label{E21}
\\ b'(T) =
\langle \log V_\mathrm{max}\rangle_T-a'\langle X\rangle_T.
\end{equation}
Different morphological types do not have identical spatial occupation,
which is shown in Fig.~\ref{F5} for Hubble types 1 and 8 as dashed
lines corresponding to forced solutions using the common slope
$a'=0.5$. The strict upper and lower cut-offs would influence the
extreme types more.
Hence we must first more carefully see if the samples
suffer from selection in $\log V_\mathrm{max}$
The inverse Tully-Fisher diagram for the diameter sample
is given in Fig.~\ref{F5}. The least squares fit
($a'=0.576$, $b'=1.259$) is shown as a solid line.
One finds evidence for both an
upper and lower cut-off in the $\log V_\mathrm{max}$-distribution,
the former being quite conspicuous.
The dotted lines are positioned
at $\log V_\mathrm{max}=2.55$ and $\log V_\mathrm{max}=1.675$
to guide the eye.
Fig.~\ref{F5} hints
that the slope $a'=0.5$ adopted in Sect.~3 may not be impeccable
and thus questions the validity of the ``na\"{\i}ve"
derivation of $H_0$ at least in the case of the diameter sample.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f6}}
\caption{The differential behaviour of $\langle\log H_0\rangle$ as
a function of the normalized distances.The inverse parameters were
$a'=0.5$ and $b'=1.450$.}
\label{F6}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f7}}
\caption{A straightforward linear regression applied to the calibrator
sample yielding $a'= 0.749$ and $b'= 1.101$.}
\label{F7}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f8}}
\caption{ As Fig.~\ref{F6}, but now the parameters
$a'= 0.749$ and $b'= 1.101$ were used.
One can see how the downward tendency between
$\log d_\mathrm{norm}\sim1.45$ and $\log d_\mathrm{norm}\sim2$ has
disappeared. Also cf. Fig. 2 in Teerikorpi et al.
(\cite{Teerikorpi99}).}
\label{F8}
\end{figure}
In the case of diameter samples,
Teerikorpi et al. (\cite{Teerikorpi99})
discussed how the cut-offs should demonstrate themselves in
a $\log H_0$ vs. $\log d_\mathrm{norm}$ diagram,
where $\log d_\mathrm{norm}=\log D_{25}+\log d_\mathrm{kin}$,
which in fact is the log of $D_\mathrm{linear}$ next to a constant.
We call $d_\mathrm{norm}$ ``normalized" in analogy to the
method of normalized distances, where the kinematical distances
were normalized in order to reveal the underlying bias. That is
exactly what is done also here.
Consider the differential behaviour of
$\langle\log H_0\rangle$ as a function of the normalized distance.
Differential average $\langle\log H_0\rangle$ was calculated as
follows. The abscissa was divided into intervals of 0.01 starting
at minimum $\log d_\mathrm{norm}$ in the sample. If a bin contained
at least 5 galaxies the average was calculated.
In Fig.~\ref{F6}. the inverse parameters $a'=0.5$ and $b'=1.450$ were
used. It is seen that around $\log d_\mathrm{n}\sim 2$ the values of
$\log H_0$ have a turning point as well as at $\log d_\mathrm{n}\sim 1.45$.
The most striking feature is -- however -- the general decreasing
tendency of $\log H_0$ between these two points.
Now, according to ET97, a downward tendency of $\log H_0$
as a function of distance corresponds to $A/A'>1$, i.e. the adopted
slope A is too shallow ($A'$ is the relevant slope).
Closer inspection of Fig.~\ref{F1} shows that a steeper slope
might provide a better fit to the calibrator sample. One is thus
tempted to ask what happens if one adopts for the field sample
the slope giving the best fit to the calibrator sample. As such
solution we adopt the straightforward linear regression yielding
$a'= 0.749$ and $b'= 1.101$ shown in Fig.~\ref{F7}.
It is interesting to note that when these parameters are used
the downward tendency between
$\log d_\mathrm{norm}\sim1.45$ and $\log d_\mathrm{norm}\sim2$
disappears as can be seen in Fig.~\ref{F8}. From hereon we refer
to this interval as the ``unbiased inverse plateau". The value
of $\log H_0$ in this plateau is still rather high.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f9}}
\caption{
The differential $\langle \log H_0\rangle$ vs. $\mu_\mathrm{norm}$
diagram. One finds no indication of a selection in
$\log V_\mathrm{max}$. The adopted slope ($a'=-0.10$) appears
to be incorrect.
}
\label{F9}
\end{figure}
In the case of the magnitude sample we study the behaviour
of the differential average $\langle \log H_0\rangle$ as
a function of a ``normalized" distance modulus:
\begin{equation}
\label{E22}
\\ \mu_\mathrm{norm} = B_\mathrm{T}^\mathrm{c}-5\log d_\mathrm{kin}.
\end{equation}
The $\mu_\mathrm{norm}$ axis was divided into intervals of 0.05
and again, if in a bin is more than five points the average is
calculated. As suspected in the view of Fig.~\ref{E4}.,
Fig.~\ref{E9} reveals no significant indications of a selection in
$\log V_\mathrm{max}$. The points follow quite well the straight
line also shown. The line however is tilted telling us that
the input slope $a'=-0.10$ may not be the relevant one.
As already noted the type corrections may have some influence
on the slopes.
In the next section we derive the appropriate type corrections
for the zero-points using galaxies residing in the unbiased
plateau ($\log d_\mathrm{norm} \in [1.45,2.0]$) for the diameter
sample and for the whole magnitude sample and rederive the
slopes.
\section{Type corrections and the value of $H_0$}
The zero-points needed for the type corrections are calculated
using Eq.~\ref{E21}. It was pointed out in Sect.~2.2 that $\log H_0$
should run horizontally in order to find an unbiased estimate for
$H_0$. In this section we look for such an slope.
Because the type-corrections depend on the adopted slope, this
fine-tuning of the slope must be carried out in an iterative manner.
This process consists of finding the type corrections
$\Delta b'(\mathrm{T})$ for each test slope $a'$. Corrections are
made for both the field and calibrator samples. The process is
repeated until a horizontal $\langle\log H_0\rangle$ run is
found.
Consider first the diameter sample. When the criteria for the
unbiased inverse plateau were imposed on the
sample, 2142 galaxies were left. For this subsample
the iteration yielded $a'=0.54$ (the straight line in
Fig.~\ref{F10} is the least squares fit with a slope 0.003)
and when the corresponding
type corrections given in Table~1 were applied to the calibrator
sample and the slope forced to it one found
$b'_\mathrm{cal}(6)=1.325$.
The result is shown in Fig.~\ref{F10}.
The given inverse parameters predict an average
$\langle\log H_0\rangle=1.897$
(or $H_0=78.9\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$).
We treated the magnitude sample of 1713 galaxies in a similar
fashion. The resulting best fit is shown in Fig.~\ref{F11}.
The relevant slope is $a'=-0.115$ (the least squares fit yields
a slope 0.0004). The corresponding type corrections are given
in Table~1. The forced calibration gives $b'_\mathrm{cal}(6)=-0.235$.
From this sample we find an average
$\langle\log H_0\rangle=1.869$
(or $H_0=72.4\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$).
In both cases the inverse estimates
for the Hubble constant ($H_0\approx80$ for the diameter relation
and $H_0\approx70$ for the magnitude relation) are considerably
larger than the corresponding estimates using
the direct Tully-Fisher relation ($H_0\approx55$).
\begin{table}
\label{T1}
\begin{center}
\begin{tabular}{ccc}
\hline
$\Delta b'(T)$&$a'=0.54$&$a'=-0.115$\\
\hline
$\Delta b'(1)$&0.125&0.110\\
$\Delta b'(2)$&0.156&0.124\\
$\Delta b'(3)$&0.129&0.096\\
$\Delta b'(4)$&0.095&0.058\\
$\Delta b'(5)$&0.069&0.030\\
$\Delta b'(6)$&0.0&0.0\\
$\Delta b'(7)$&-0.054&-0.042\\
$\Delta b'(8)$&-0.118&-0.075\\
\hline
\end{tabular}
\end{center}
\caption{The type corrections required for the relevant
slopes $a'=0.54$ for the unbiased diameter sample and $a'=-0.115$
for the magnitude sample.}
\end{table}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f10}}
\caption{The differential $\langle\log H_0\rangle$
as a function of the log of normalized
distance $\log d_\mathrm{norm}$
for the plateau galaxies with the adopted relation
$\log V_\mathrm{max}=0.54\log D+1.325$. The solid line
is the average $\log H_0=1.897$.}
\label{F10}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f11}}
\caption{The differential $\langle\log H_0\rangle$
as a function of the log of normalized
distance modulus $\mu_\mathrm{norm}$
for the plateau galaxies with the adopted relation
$\log V_\mathrm{max}=-0.115M-0.235$. The solid line is
the average $\log H_0=1.869$.}
\label{F11}
\end{figure}
\section{$H_0$ corrected for a calibrator selection bias}
The values of $H_0$ from the direct and inverse relations
still disagree {\it even} after we have taken into account
the selection in $\log V_\mathrm{max}$, made the type corrections
and used the relevant slope.
There is -- however -- a serious possibility
left to explain the discrepancy.
{\it The calibrator
sample used may not meet the theoretical requirements
of the inverse relation}.
In order to transform the relative distance scale
into an absolute one a properly chosen sample
of calibrating galaxies is needed.
What does ``properly chosen" mean?
Consider first the direct relation for which
it is essential to possess a
calibrator sample, which is volume-limited for {\it each}
$p_\mathrm{cal}$. This means that for a $p_\mathrm{cal}$
one has $X_\mathrm{cal}$
which is drawn from the complete part of the gaussian distribution
function $G(X;X_p,\sigma_{X_p})$, where the average $X_p=ap+b$.
If $\sigma_{X_p}$ is constant for all $p$ and the
direct slope $a$ has been {\it correctly} derived
from the unbiased field
sample, it will, when forced onto the calibrator sample,
bring about the correct calibrating zero-point.
As regards the calibration of the inverse relation the
sample mentioned above does not necessarily guarantee
a successful calibration. As pointed out by Teerikorpi
et al. (\cite{Teerikorpi99}) though the calibrator
sample is complete in the {\it direct} sense nothing has
been said about how the $p_\mathrm{cal}$'s relate to the
corresponding cosmic distribution of $p$'s from which the
field sample was drawn. $\langle p\rangle_\mathrm{cal}$
should reflect the cosmic average $p_0$.
If not, the relevant field slope when forced
to the calibrator sample will bring about a biased estimate
for $H_0$.
Teerikorpi (\cite{Teerikorpi90})
already recognized that this
could be a serious problem. He studied, however,
clusters of galaxies where a nearby (calibrator) cluster
obeys a different slope than a distant cluster.
Teerikorpi et al. (\cite{Teerikorpi99})
developed the ideas further
and showed how this problem may be met also when using
field galaxies. The mentioned bias when using the relevant
slope can be corrected
for but is a rather complicated task. For the theoretical
background of the ``calibrator selection bias"
consult Teerikorpi et al. (\cite{Teerikorpi99}).
One may -- as pointed out by Teerikorpi et al.
(\cite{Teerikorpi99}) -- use instead of the relevant
slope the calibrator slope which also predicts a biased
estimate for $H_0$ but which can be corrected for in
a rather straightforward manner.
For the diameter relation the average correction
term reads as
\begin{equation}
\label{E23}
\\ \Delta\log H_0=(3-\alpha)\,\ln 10\,\sigma_D^2
\times\left[\frac{a'_\mathrm{cal}}{a'}-1\right],
\end{equation}
where $\sigma_D$ is the dispersion of the log linear
diameter $\log D_\mathrm{linear}$ and $\alpha$ gives
the radial number density gradient : $\alpha=0$ corresponds to
a strictly homogeneous distribution of galaxies.
For magnitudes the correction term follows from
(cf. Teerikorpi \cite{Teerikorpi90})
\begin{equation}
\label{E24}
\\ \Delta\log H_0=
0.2\,\left[\frac{a'_\mathrm{cal}}{a'}-1\right]
\times(\langle M\rangle-M_0).
\end{equation}
Because $\langle M\rangle-M_0$ simply reflects the classical
Malmquist bias one finds:
\begin{equation}
\label{E25}
\\ \Delta\log H_0=\frac{(3-\alpha)\ln 10}{5}\,\sigma_M^2
\times0.2\left[\frac{a'_\mathrm{cal}}{a'}-1\right],
\end{equation}
Note that one may use the calibrator slope and consequently
the correction formulas {\it irrespective} of the nature
of the calibrator sample
(Teerikorpi et al. \cite{Teerikorpi99}). If the calibrator
sample would meet the requirement mentioned, the value corrected with
Eqs.~\ref{E23} or~\ref{E25} should equal values obtained
from the relevant slopes. Furthermore, our analysis carried
out so far would have yielded an unbiased estimate for $H_0$
and thus the problems would be in the direct analysis.
However, if the requirement is not met one should prefer the
corrective method using the calibrator slope.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f12}}
\caption{Histogram of the
$\log V_\mathrm{max}$ values and the
individual calibrators (labelled with stars).
The vertical solid line gives the median of
the plateau
$\mathrm{Med}(\log V_\mathrm{max}^\mathrm{plateau})=2.10$
and the dotted line gives the median of the calibrators
$\mathrm{Med}(\log V_\mathrm{max}^\mathrm{calib})=2.11$.}
\label{F12}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f13}}
\caption{The Kolmogorov-Smirnov test for the diameter sample.
Pay attention to the rather remarkable similarity between
the cumulative distribution functions (cdfs).}
\label{F13}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f14}}
\caption{The Kolmogorov-Smirnov test for the magnitude sample. Again
the cdfs are quite similar.}
\label{F14}
\end{figure}
\subsection{Is the calibrator sample representative?}
Is the calibrator bias present in our case? Recall that the
calibrators used were sampled from the nearby field
to have high quality distance moduli mostly from the
HST Cepheid measurements.
This means that we have no a priori guarantee that the
calibrator sample used will meet the criterium
required.
We compare the type-corrected diameter and magnitude
samples with the calibrator sample. Note that for the
diameter sample we use only galaxies residing in the
unbiased inverse plateau (i.e. the small selection effect
in $\log V_\mathrm{max}$ has been eliminated).
In Fig.~\ref{F12} we show the histogram of the
$\log V_\mathrm{max}$ values for the diameter sample
and the individual calibrators (labelled as stars).
The vertical solid line gives the median of
the plateau
$\mathrm{Med}(\log V_\mathrm{max}^\mathrm{plateau})=2.10$
and the dotted line gives the median of the calibrators
$\mathrm{Med}(\log V_\mathrm{max}^\mathrm{calib})=2.11$.
In the case of magnitudes both the field and calibrator
sample have the same median (2.14).
The average values for the diameter case were
$\langle\log V_\mathrm{max}^\mathrm{plateau}\rangle=2.09$
and
$\langle\log V_\mathrm{max}^\mathrm{calib}\rangle=2.06$,
and for the magnitude case
$\langle\log V_\mathrm{max}^\mathrm{mag}\rangle=2.12$
and
$\langle\log V_\mathrm{max}^\mathrm{calib}\rangle=2.08$.
Both the diameter and the magnitude field samples were
subjected to strict limits, which means that both
inevitably suffer from the classical Malmquist bias.
In order to have a representative calibrator sample in the sense
described, we would have expected a clear difference
between the field and calibrator samples. That the statistics
are very close to each other lends credence to the assumption
that the calibrator selection bias is present.
We also made tests using the Kolmogorov-Smirnov
statistics (Figs.~\ref{F13} and~\ref{F14}).
In this test a low significance level
should be considered as counterevidence for a hypothesis that two
samples rise from the same underlying distribution.
We found relatively high significance
levels (0.89 for the diameter sample and 0.3 for the magnitude sample).
Neither these findings corroborate the
hypothesis that the calibrator sample is drawn from the cosmic
distribution and hence the use of Eqs.~\ref{E23} or~\ref{E25}
is warranted.
\subsection{The dispersion in $\log D_\mathrm{linear}$}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f15}}
\caption{A classical Spaenhauer diagram for normalized distances vs.
kinematical distances with a presumed dispersion $\sigma_X=0.28$.
}
\label{F15}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f16}}
\caption{Comparison between average values of
$\langle\log d_\mathrm{norm}\rangle$
for different kinematical distances and the theoretical prediction
calculated from Eq.~\ref{E26} with $X_0^*=1.37$ and $\sigma_X=0.28$.
}
\label{F16}
\end{figure}
In order to find a working value for the dispersion in
$\log D_\mathrm{linear}$, we first consider the classical
Spaenhauer diagram (cf. Sandage \cite{Sandage94a},
\cite{Sandage94b}).
In the Spaenhauer diagram one studies the behaviour of
$X$ as a function of the redshift. If the observed
redshift could be translated into the corresponding cosmological distance,
then $X$ inferred from $x$ and the redshift would genuinely
reflect the true size of a galaxy.
In practice, the observed redshift cannot be considered as
a direct indicator of the cosmological distance because of the
inhomogeneity of the Local Universe. Peculiar motions
should also be considered. Thus the inferred $X$
suffers from uncertainties in the underlying kinematical model.
The Spaenhauer diagram as a diagnostics for the distribution function
is always constrained by our knowledge of the form of the true
velocity-distance law.
Because the normalized distance (cf. Sect.~3.) is proportional
to the linear diameter we construct the Spaenhauer diagram as
$\log d_\mathrm{norm}$ vs. $\log d_\mathrm{kin}$ thus avoiding the
uncertainties in the absolute distance scale. The problems with
relative distance scale are -- of course -- still present.
The fit shown in Fig.~\ref{F15} is not unacceptable.
The dispersion used was $\sigma_X=0.28$, a value inferred
from the dispersion in absolute B-band magnitudes $\sigma_M=1.4$
(Fouqu\'e et al. \cite{Fouque90})
based on the expectation that the
dispersion in log linear diameter should be one fifth of that
of absolute magnitudes.
We also looked how the average values
$\langle\log d_\mathrm{norm}\rangle$
at different kinematical distances compare to the theoretical
prediction which, in a strictly limited sample of $X$'s,
at each log distance is formally expressed as
\begin{equation}
\label{E26}
\\ \langle X\rangle_\mathrm{d} = X_0^*+
\frac {2\sigma_X}{\sqrt2\pi}
\frac
{\exp\bigl[-(X_\mathrm{lim}-X_0^*)^2/(2\sigma_X^2)\bigr]}
{\mathrm{erfc}\bigl[(X_\mathrm{lim}-X_0^*)/(\sqrt2\sigma_X)\bigr]}.
\end{equation}
Here $X$ refers to $\log d_\mathrm{norm}$.
The curve in Fig.~\ref{F16} is based on $X_0^*=1.37$
and $\sigma_X=0.28$. The averages from the data are shown as bullets.
The data points follow the theoretical prediction reasonably well.
\subsection{Corrections and the value of $H_0$}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f17}}
\caption{A least squares fit the type corrected calibrator
sample yielding $a'=0.73$.
The type correction was based on $a'=0.54$.}
\label{F17}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics[211,195][460,384]{8156.f18}}
\caption{A least squares fit the type corrected calibrator
sample yielding $a'=-0.147$.
The type correction was based on $a'=-0.115$.}
\label{F18}
\end{figure}
Consider a strictly homogeneous universe, i.e. $\alpha=0$.
In Eqs. \ref{E23} and \ref{E25} one needs
values for slope $a'_\mathrm{c}$.
Least squares fit to the type-corrected calibrator sample yields
$a'_\mathrm{c}=0.73$ for the diameter relation
and $a'_\mathrm{c}=-0.147$ for the magnitude relation.
(cf. Figs. \ref{F17} and \ref{F18}).
These slopes correspond to diameter zero-point
$b'_\mathrm{c}(6)=1.066\pm0.103$
and to magnitude zero-point
$b'_\mathrm{c}=-0.879\pm0.131$
The {\it biased} estimates for average $\log H_0$ are
$\langle\log H_0\rangle = 1.910\pm0.188$ for the diameters
and $\langle\log H_0\rangle = 1.876\pm0.176$ for the magnitudes.
For the zero-points and the averages we have given the $1\sigma$
standard deviations. The {\it mean error} in the averages is
estimated from
\begin{equation}
\label{E27}
\\ \epsilon_{\langle\log H_0\rangle}\approx
\sqrt{\frac{\sigma_{B'}^2}{N_\mathrm{cal}}
+\frac{\sigma_{\log H_0}^2}{N_\mathrm{gal}}},
\end{equation}
where $\sigma_{B'}=\sigma_{b'}/a'_\mathrm{cal}$ for
diameters and $\sigma_{B'}=0.2\sigma_{b'}/a'_\mathrm{cal}$
for magnitudes. The use of Eq.~\ref{E27} is acceptable
because the dispersion in $b'$ does not correlate with
the dispersion $\log H_0$. With the given slopes and dispersions
we find:
\begin{itemize}
\item $\langle\log H_0\rangle = 1.910\pm0.037$ for the diameters
\item $\langle\log H_0\rangle = 1.876\pm0.046$ for the magnitudes.
\end{itemize}
Eq. \ref{E23} predicts an average correction term for
the slopes $a'_\mathrm{c}=0.73$ and $a'=0.54$
together with $\sigma_X=0.28$ $\Delta\log H_0=0.191$.
and Eq. \ref{E25} with $a'_\mathrm{c}=-0.147$,$a'=-0.115$
and $\sigma_M=1.4$ $\Delta\log H_0=0.151$. When applied
to the above values we get the corrected, unbiased estimates
\begin{itemize}
\item $\langle\log H_0\rangle= 1.719\pm0.037$ for the diameters
\item $\langle\log H_0\rangle = 1.725\pm0.046$ for the magnitudes.
\end{itemize}
These values translate into Hubble constants
\begin{itemize}
\item $H_0=52^{+5}_{-4}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse diameter
B-band Tully-Fisher relation, and
\item $H_0=53^{+6}_{-5}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse magnitude
B-band Tully-Fisher relation.
\end{itemize}
These corrected values are in good concordance with each
other as well as with the estimates established from
the direct diameter Tully-Fisher relation
(Theureau et al. \cite{Theureau97b}).
Note that the errors in the magnitude relation are slightly
larger than in the diameter relation. This is expected because
for the diameter relation we possess more galaxies. The error
is however mainly governed by the uncertainty in the calibrated
zero-point. This is expected because though the dispersion in
inverse relation as such is large it is compensated by the
number galaxies available.
Finally, how significant an error do the correction formulae
induce? We suspect the error to mainly depend on $\alpha$.
The correction above was based on the assumption of
homogeneity (i.e. $\alpha=0$). Recently
Teerikorpi et al. (\cite{Teerikorpi98}) found
evidence that the average density radially
decreases around us ($\alpha\approx0.8$)
confirming the more general (fractal) analysis
by Di Nella et al. (\cite{DiNella96}).
Using this value of $\alpha$
we find $\Delta\log H_0=0.140$ for the diameters
and $\Delta\log H_0=0.111$ for the magnitudes
yielding
\begin{itemize}
\item $\langle\log H_0\rangle = 1.770\pm0.037$ for the diameters
\item $\langle\log H_0\rangle = 1.765\pm0.046$ for the magnitudes.
\end{itemize}
In terms of the Hubble constant we find
\begin{itemize}
\item $H_0=59^{+5}_{-4}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse diameter
B-band Tully-Fisher relation, and
\item $H_0=58^{+6}_{-5}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse magnitude
B-band Tully-Fisher relation.
\end{itemize}
\section{Summary}
In the present paper we have examined how to apply
the inverse Tully-Fisher relation to the problem of
determining the value of the Hubble constant,
$H_0$, in the practical context of the large galaxy sample KLUN.
We found out that the implementation
of the inverse relation is not as simple task as one
might expect from the general considerations (in particular the quite
famous result of the unbiased nature of the relation).
We summarize our main results as follows.
\begin{enumerate}
\item A straightforward application of the inverse relation
consists of finding the average Hubble ratio for each kinematical
distance and tranforming the relative distance into an absolute one
through calibration. The 15 calibrator galaxies used
were drawn from the field
with cepheid distance moduli obtained mostly from the HST
observations. The inverse diameter relation predicted
$H_0\approx 80\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
and the magnitude relation predicted
$H_0\approx 70\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
The diameter value for $H_0$ is about 50 percent and
the magnitude value about 30 percent larger than those
obtained from the direct relation
(cf. Theureau et al. \cite{Theureau97b}).
\item We examined whether this discrepancy could be resolved
in terms of some selection effect in $\log V_\mathrm{max}$
and the type dependence of the zero-points on the Hubble type.
One expects these to have some influence on the
derived value of $H_0$. Only a minuscule effect was observed.
\item There is -- however -- a new kind of bias involved:
if the $\log V_\mathrm{max}$-distribution
of the calibrators does not reflect the cosmic distribution
of the field sample {\it and} the relevant slope for the field galaxies
differs from the calibrator slope the average value of $\log H_0$
will be biased if the relevant slope is used
(Teerikorpi et al. \cite{Teerikorpi99}).
\item We showed for the unbiased inverse plateau galaxies
i.e. a sample without galaxies probably suffering from selection in
$\log V_\mathrm{max}$,
that the calibrators and the field sample obey different inverse
diameter slopes, namely $a'_\mathrm{cal}=0.73$ and $a'= 0.54$,
Also, the magnitude slopes differed from each other
($a'_\mathrm{cal}=-0.147$ and $a'= -0.115$). For the
diameter relation we were able to use 2142 galaxies and for
the magnitude relation 1713 galaxies. These sizes are
significant.
\item We also found evidence that the calibrator sample
does {\it not} follow the cosmic distribution of
$\log V_\mathrm{max}$ for the field galaxies. This means
that if the relevant slopes are used a too large value for $H_0$
is found. Formally, this calibrator selection bias could be
corrected for but is a complicated task.
\item One may use instead of the relevant slope the calibrator
slope which also brings about a biased value of $H_0$. Now,
however, the correction for the bias is an easy task.
Furthermore, this approach can be used irrespective of the nature
of the calibrator sample and should yield an unbiased estimate
for $H_0$.
\item When we adopted this line of approach we found
\begin{itemize}
\item $H_0=52^{+5}_{-4}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse diameter
B-band Tully-Fisher relation, and
\item $H_0=53^{+6}_{-5}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse magnitude
B-band Tully-Fisher relation
\end{itemize}
for a strictly homogeneous distribution of galaxies ($\alpha=0$) and
\begin{itemize}
\item $H_0=59^{+5}_{-4}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse diameter
B-band Tully-Fisher relation, and
\item $H_0=58^{+6}_{-5}\mathrm{\ km\,s^{-1}\,Mpc^{-1}}$
for the inverse magnitude
B-band Tully-Fisher relation.
\end{itemize}
for a decreasing radial density
gradient ($\alpha=0.8$).
\end{enumerate}
These values are in good concordance with each other as well as
with the values established from the corresponding direct
Tully-Fisher relations derived by Theureau et al. (\cite{Theureau97b}),
who gave a strong case for
the long cosmological distance scale consistently supported by
Sandage and his collaborators. Our analysis also
establishes a case supporting such a scale.
It is worth noting that this is the first time when the
{\it inverse} Tully-Fisher relation clearly lends credence
to small values of the Hubble constant $H_0$.
\begin{acknowledgements}
{We have made use of data from the Lyon-Meudon extragalactic
Database (LEDA) compiled by the LEDA team at the CRAL-Observatoire
de Lyon (France). This work has been supported by the Academy of Finland
(projects ``Cosmology in the Local Galaxy Universe'' and
``Galaxy streams and structures in the Local Universe'').
T. E. would like to thank G. Paturel and his staff for hospitality
during his stay at the Observatory of Lyon in May 1998. Finally,
we thank the referee for useful comments and constructive
criticism.}
\end{acknowledgements}
|
\section{Introduction}
Planetary nebulae (PNe) are extraordinarily useful for probing stellar
evolution and cosmology. In extragalactic astronomy, the planetary-nebula
luminosity function (PNLF) is one of the most accurate and reliable indicators
of relative distance (see the review of Jacoby {\refit et~al.\ } 1992); in stellar
astrophysics, PNe allow us to examine the physics of mass loss and the
timescales of stellar evolution (e.g., the review of Iben 1995). Since young
PNe are bright emission-line sources, they make ideal test particles for
dynamical studies, both in the Milky Way and in external galaxies ({\refit e.g.,}\
Ciardullo, Jacoby, \& Dejonghe 1993; Amaral {\refit et~al.\ } 1996). Finally, since the
chemical abundances in a PN reflect the chemistry of the interstellar medium
at the time of its progenitor's formation (with some light species possibly
altered subsequently by stellar nucleosynthesis), these objects can yield
unique insights into galactic star-formation histories and chemical and
stellar evolution (cf.~Dopita {\refit et~al.\ } 1997).
Almost every interesting quantity related to PNe in the Milky Way---their
space density, formation rate, Galactic distribution, sizes, ionized and
total nebular masses, contribution to chemical evolution, and the luminosities
and evolutionary states of their central stars---depends critically upon their
distances. But, unfortunately, the distances to PNe within the Galaxy are
known only poorly. It is a remarkable irony that while the PNLF can be used to
derive relative distances to external galaxies to better than $\sim 10\%$
(cf.~Jacoby {\refit et~al.\ } 1992), the distances to Milky Way PNe are typically known to
no better than a factor of $\sim 2$ or worse ({\refit e.g.,}\ Terzian 1993, 1997). In
fact, of the $\sim 1100$ known Galactic PNe, only a dozen or so have distances
that are reasonably well determined using direct methods. For the rest, it is
necessary to use various statistical techniques.
One fundamental, but rarely used, method for obtaining distances to Galactic
PNe is through the photometric parallaxes of resolved companion stars.
Perhaps two-thirds of all stars are members of binary systems, and the
evidence suggests that the orbital period distribution of these binaries is a
Gaussian in $\log P$ centered at $P \approx 180$~years, with a dispersion of
2.3 in the logarithm of the period (Duquennoy \& Mayor 1991). Main-sequence
binaries with periods less than $\sim 1000$~days, but still wide enough for a
red giant to form, will eventually evolve through a common-envelope phase and
produce systems with dramatically shorter periods or even a coalesced binary
(see Bond \& Livio 1990; Yungelson, Tutukov, \& Livio 1993). However, stars
with larger initial periods will not interact, and their separations will
actually increase with time, as a consequence of stellar mass loss. As a
result, it is likely that nearly half of all planetary-nebula nuclei (PNNs)
have wide binary companions.
Despite this expectation, only a few PNNs are actually known to have resolved
visual companions. The best-known case is the nucleus of NGC~246, which has a
14th-mag K-dwarf companion $3\farcs 8$ away (Minkowski 1960; Cudworth 1973).
Fitting of this companion to the main sequence provides what is generally
accepted as one of the most accurate PN distances (quoted as 430~pc by Cahn,
Kaler, \& Stanghellini 1992, and revised to $495^{+145}_{-100}$~pc on the
basis of CCD photometry by Bond \& Ciardullo 1999), and one that is almost
always used as a primary calibrator of statistical distance methods. Other
PNNs reported to have resolved companions include those of NGC 650-1, A~24,
A~30, and A~33 (Cudworth 1973), NGC~3132 (Kohoutek \& Laustsen 1977), NGC~6853
(Cudworth 1973, 1977), A~63 (Krzeminski 1976), K~1-14 (Kaler 1981), and PuWe~1
(Purgathofer \& Weinberger 1980), but at the time of this survey, only the
companion to NGC~3132 had photometry accurate enough to provide a reliable
spectroscopic parallax (Pottasch 1980, 1984).
Here we present the results of a {\sl Hubble Space Telescope\/} ({\sl HST\/}\/)
snapshot survey of Galactic planetary nebulae, designed to detect and measure
resolved binary PNN companions. In \S 2, we describe the survey and
the photometric procedures used to measure all the stars present on our CCD
frames. In \S 3, we use these data to create a list of PNNs which have
visual companions deserving of follow-up observations. In \S 4, we discuss
the issue of interstellar reddening, and compare measures of extinction
based on two-color photometry with estimates derived from the emission lines
of the nebulae. In \S 5, we transform our {\sl HST\/}\/ stellar magnitudes to the
standard Landolt (1983, 1992) system, and present $V$ and $I$ magnitudes for
109 central stars. Because of the superior resolving power of the {\sl HST\/},
these magnitudes are generally better than PNN measurements made from the
ground, especially for objects with $V > 15.4$. In \S 6, we discuss our
candidate binaries individually, and derive distances to those
systems that are most likely to be physically associated. Finally, in \S 7, we
compare our distances to estimates based on statistical methods, and discuss
the implications our observations have for the Galactic PN distance scale.
\section{Target Selection, Observations, and Reductions}
Our observations were carried out as a ``snapshot survey'' during Cycles~3
and~5 of the {\sl HST\/}\/ General Observer program. {\sl HST\/}\/ snapshots are short
exposures taken during occasional gaps that remain in the observing program
after as many primary scientific observations as possible have been scheduled.
The targets that are observed during these opportunities are selected at
random from a list of candidates provided by the observers.
The PNe in our target pool were chosen using criteria designed to maximize our
chances of finding resolved companion stars. First among these criteria was
that the objects in our survey had to be nearby. Our target selection was
heavily influenced by the list of very nearby PNe given by Terzian (1993), and
nearly all of our targets have statistical distances from Cahn, Kaler, \&
Stanghellini (1992; hereafter CKS) and/or Zhang (1995) that are less than
$\sim 3$~kpc. Almost as important was our Galactic-latitude criterion: in
order to reduce contamination from superposed field stars, most of our targets
were chosen to have $|b| > 10^\circ$. A third selection criterion was known
or suspected binarity of the central star. We included in our target list the
nine PNNs with reported visual companions (as listed above), along with A~34,
A~66, and Th~2-A, whose visual companions were noted by H.E.B. during
ground-based observations. In addition we included seven central stars that
have composite or late-type spectra but are unresolved from the ground, and
show no evidence for being extremely close binaries. Finally, to increase the
usefulness of our dataset for comparison with other distance techniques, we
included six PNe with Very Large Array expansion distances (Hajian, Terzian,
\& Bignell 1993, 1995; Hajian \& Terzian 1996), 15 nebulae with distance
estimates based on model-atmosphere analyses of their central stars (M\'endez,
Kudritzki, \& Herrero 1992), and seven PNNs known to be very close binaries
(A~46, A~63, A~65, HFG~1, K~1-2, LoTr~5, and NGC~2346; cf.~Bond 1994), from
which distances can potentially be obtained from light-curve solutions. In
all, 144~PNNs were included in our input target lists.
Our snapshot images were taken with the {\sl HST\/}\/ between 1993 and 1997.
The Cycle~3 (1993) observations were obtained with the Planetary Camera
of the original Wide Field and Planetary Camera (WF/PC){}. These data
consisted of single exposures through the F785LP (``$I$'') filter plus
occasional exposures through the F555W (``$V$'') filter, usually with the
PN central star positioned near the center of the PC6 CCD{}.
The Cycle~5
(1995-97) data were obtained using the F814W (``$I$'') and F555W
(``$V$'') filters of the Wide Field Planetary Camera~2 (WFPC2), with the
CCD gain set at 14 e$^-$/DN{}. For these observations the PNNs were centered
on the Planetary Camera chip and usually imaged twice in each filter, with
the second exposure typically being $\sim 2$ times longer than the first.
The uneven exposure times optimized our ability to detect binaries.
We generally tried to scale our shorter integrations so that the PNN's image
fell just short of saturation ({\refit i.e.,}\ $\sim 50,000$~electrons in the central
pixel); this insured an accurate measurement of the PNN's magnitude and color,
and enabled us to search for companions with $\Delta m \lesssim 4.5$ to
within $\rho\sim 0\farcs 05$ of the central star. Conversely, we attempted to
scale our longer exposures so that the central pixel of the PNN could receive
as much charge as possible ($\sim 100,000$~electrons) without bleeding
significantly into its neighboring pixels. By doing this, we could search
for companions almost $\sim 7$~mag fainter than the central star with
$\rho\gtrsim 0\farcs 3$ separation.
Because of the limited time available per snapshot visit, we imposed a maximum
integration time of 900~s per filter. The second, deeper exposures in each
filter described above were omitted if they would be longer than 500~s in
F814W, or 200~s in F555W{}. Exposures longer than 600~s were split into two
equal-length integrations to aid in cosmic-ray removal. Because of the
uncertainties of the ground-based PNN magnitudes used to calculate our
exposure times, we had anticipated that not all of our targets would be
optimally exposed. In fact, however, most of our program PNNs had useful
exposures, and we were able to perform a careful search for visual binaries
around almost every object. There was a guide-star acquisition error for our
observation of Lo~4, which resulted in our obtaining only a single F555W
exposure under gyro control.
A summary of the WF/PC and WFPC2 observations appears in Tables~1 and 2. Of
the 144 objects in the target list, 113 were actually observed: 26 with WF/PC,
84 with WFPC2, and 3 with both.
Photometric reduction of our CCD images was accomplished using a combination
of IRAF and DAOPHOT/ALLSTAR (Stetson 1987). After the standard STScI pipeline
de-biasing and flat-fielding, we removed charged-particle events (``cosmic
rays'') from the data. For this task, the Cycle~3 and Cycle~5 frames
were handled differently. Because the Cycle~3 data consisted only of
single exposures through the F785LP and F555W filters, we could not
simply ``stack'' these images and remove discrepant pixels. Instead
we employed a semi-automatic iterative technique, which took advantage
of the fact that stars on the original WF/PC frames had aberrated
point-spread-functions (PSFs). First, we filtered our images using a
$7 \times 7$ moving window, and masked all points that deviated by more
than $3 \, \sigma$ from the median defined by their surroundings. We
then compared these filtered images to their parent images, and carefully
examined the cores of all stars on the frames. If the filtering algorithm
had affected the cores, the original pixel values were manually inserted
back into the images. In this way, the bulk of the cosmic rays were removed
without affecting the stellar photometry.
Reduction of the WFPC2 data proceeded in a more standard manner. All the data
were first multiplied by a corrective image, in order to compensate for
the geometrical distortions present in the flat-field images. We then used the
DAOPHOT routine {\it find\/} to produce a list of candidate objects for each
frame. These lists were culled of cosmic-ray events and other spurious
detections by first rejecting all objects that fell within the vignetted
areas of the frames, and then comparing the object lists derived from the
individual frames of a given target. Objects found on the shorter exposures,
but not on the longer ones, were discarded as cosmic rays.
Photometry of the WFPC2 data was performed via a PSF fitting technique.
Using all the frames of our survey, we identified $\sim 20$ bright, isolated
stars in each of the four CCD camera fields, and from these data we defined the
instrument's position-dependent PSF{}. Once the PSF was defined, we used
it to derive preliminary instrumental magnitudes and create star-subtracted
images. If any poor subtractions were found, due either to the presence
of an undetected close binary, or to a poorly determined nebular background,
the process was repeated with a modified star list and/or a model for the
nebular emission (cf.~Ciardullo \& Bond 1996). The raw instrumental
magnitudes derived by ALLSTAR were then scaled to an 11-pixel-radius aperture
using an aperture correction derived from the bright stars, and corrected for
the effects of finite charge-transfer efficiency following the prescription of
Holtzman {\refit et~al.\ } (1995). Finally, the WFPC2 magnitudes were corrected for a
non-linear term related to integration time (Casertano 1997), and put on the
photometric system defined by Holtzman {{\null} et al.}
Because the aberrated WF/PC images had a brighter limiting magnitude than
their WFPC2 counterparts, there were, in general, fewer stars on these frames
and saturation in the stellar cores was less of a problem. Thus, our
photometric techniques were simpler. In most cases,
photometry was accomplished by summing the flux of each star within an
aperture of radius 5~pixels; only when the stellar separations were of the
order of a few pixels, or when the stellar cores were saturated, were
DAOPHOT's PSF fitting routines used. The small-aperture measurements were then
converted to total instrumental magnitudes using aperture corrections found
from bright stars, and scaled to the {\sl HST\/}\/ standard system using the
ground-based WF/PC calibration of Harris {\refit et~al.\ } (1991) and the flight
calibration of Hunter {\refit et~al.\ } (1992). Finally, depending on the Julian Date of
the observation, an additional correction of up to $0.1$~mag was applied to
the F555W magnitudes, in order to account for the WF/PC chips' loss of
sensitivity with time due to contamination (cf.~Ritchie \& MacKenty 1993,
1994).
\section{Detection of Binaries}
Because our data contain no information about proper motion or radial
velocity, our assignment of physical binarity has to rely solely on the
spatial coincidence of stars. The probability of our identifying a binary PNN
depends critically both on the apparent separation of the companion star, and
on the field-star surface density. (In other words, in a sufficiently crowded
field even a closely separated physical pair would be mistaken for an optical
double.) To improve our chances of detecting visual binaries, our target
sample was weighted toward objects at high galactic latitude and toward nearby
objects, as described above. Still, because the surface density of bright
stars is much lower than for faint stars, our analysis is significantly more
sensitive to brighter companions than to fainter ones.
To estimate whether a given apparent companion star located near a PNN is
physically associated with it, we calculate the Poissonian probability $P$
that a random field star as bright as, or brighter than, the companion would
be projected by chance within a radius $\rho$ of the PNN{}. Mathematically,
\begin{equation}
P = 1 - \left(1 - {\pi \rho^2 \over A} \right)^N,
\end{equation}
where $A$ is the total sky area surveyed by each WF/PC or WFPC2 frame, and $N$
is the number of stars within this area that are at least as bright as the
companion. $P$ is a function of the local stellar surface density,
which we determined directly from star counts on the frames.
If true physical binaries are common in our sample, there should be an excess
of pairs with very low values of $P${}. (It was, of course, just such an
argument, first made by Michell 1767, that established the existence of
physical double stars.) In Figure 1 we plot the distribution of $P$ for our
sample of PNNs, using the stellar neighbor of each central star that has the
lowest value of $P$ (i.e., the highest probability of being physically
associated; generally, but not always, this is also the nearest neighbor of
the PNN){}. If all of our frames were populated entirely by randomly
distributed field stars, without any true physical pairs, the distribution
would be flat; instead, Figure~1 shows a large excess of companions with very
small chance probabilities, particularly for $P\le0.05$. This is a clear
demonstration that we are detecting true resolved physical doubles.
We therefore chose all of the apparent binaries with a chance probability of
5\% or less for further investigation, and list them in Table~3. Most of the
columns are self-explanatory. Columns~2 and 3 give the separation, $\rho$,
between the PNN and the candidate companion, and the J2000 position angle of
the companion with respect to the central star. Column~8 presents the
probability $P$ of a chance projection, as calculated from equation~(1). To
give an indication of the uncertainty in the probability, we give in the final
column of Table~3 the value of $P$ calculated by increasing the local stellar
density by $1 \, \sigma$ above that deduced directly from the star counts.
Table~3 is {\it not\/} necessarily a list of true binary PNNs: it is merely
the complete set of objects that, formally, have more than a 95\% probability
of being a physical pair. Since we imaged over a hundred PNNs in the course of
this survey, we can expect $\sim 6$ chance superpositions to be contained in
the list. In \S 6 we will examine the properties of the PNNs in question and
attempt to select those systems that are most likely to be true physical
pairs.
Returning to Figure~1, we note that in addition to the sharp peak at small
probabilities, there is a slight excess of objects with chance probabilities
between 0.05 and 0.40. This suggests that a few additional PNNs not included
in Table~3 also have physical companions. Unfortunately, because the ratio of
true binaries to chance superpositions in this probability range is low, the
only way to identify these objects would be through proper-motion and/or
radial-velocity measurements of the candidate companions.
\section{Interstellar Extinction}
The first step in deriving distances from two-color photometry is to obtain
an estimate of the amount of interstellar extinction suffered by each object.
This can be done in two ways.
The first method is to use the observed emission ratios ({\refit i.e.,}\ the relative
intensities of H$\alpha$, H$\beta$, and the radio continuum) in the PNNs'
surrounding nebulae. The intrinsic {H$\alpha$\ } to {H$\beta$\ } ratio of a typical
$10,000$~K nebula is $\sim 2.86$ ({\refit e.g.,}\ Brocklehurst 1971); by comparing this
value to the observed line ratio, it is possible to compute the logarithmic
extinction at H$\beta$, denoted $c$. Similarly, an estimate of $c$ can be
obtained by comparing the nebula's radio continuum (measured at 5~GHz) to its
total emission at H$\beta$ (cf.~Milne \& Aller 1975). Measures of $c$ exist
in the literature for a large number of objects, and a useful compilation
appears in CKS\null. Unfortunately, because most PNe exhibit complex
multi-zone structures, the exact values to use for the intrinsic
H$\alpha$/{H$\beta$\ } and H$\beta$/radio ratios can be difficult to determine; in
addition, observational errors introduced by atmospheric dispersion,
background confusion (for radio-faint PNe), and uncertain emission-line
photometry (for large and low-surface-brightness PNe) can lead one astray.
Finally, as our {\sl HST\/} images of NGC~7027 and IC~4406 in Figure~2 show,
the dust within a PN can be extremely patchy. Thus adoption of a global
extinction value for the particular line of sight to the central star (or to
its companion) may not be appropriate.
The second method of determining the extinction to a PNN is to compare the
observed color of the central star with that expected from a hot source.
Simulations using the WFPC2 efficiency curves (Biretta {\refit et~al.\ } 1996) demonstrate
that a star whose optical continuum falls on the Rayleigh-Jeans tail of the
Planck function should have an F555W$-$F814W color of $\sim -0.4$. As the
histogram of Figure~3 illustrates, this value is confirmed in our survey: the
distribution of PNN colors observed with WFPC2 has a hard blue limit of
${\rm F555W} - {\rm F814W} \sim -0.4$, and a long tail to the red caused by
interstellar (and/or circumstellar) reddening. Values for the extinction
of each PNN can therefore be obtained directly from our measured colors. Of
course, some of the same uncertainties that affect the nebula-based extinction
ratios apply here: spectroscopy demonstrates that there are real departures
from Rayleigh-Jeans continua in some objects ({\refit e.g.,}\ from Wolf-Rayet emission
lines; see, for example, Smith \& Aller 1969; Heap 1982; M\'endez {\refit et~al.\ } 1986).
Moreover, some PNNs have intrinsically red colors (either from an unresolved
late-type companion, or because they themselves have evolved back to a
``born-again'' red-giant phase). Finally, even if all central stars were
identical, small-scale non-uniformities in the dust distribution (such as seen
in Figure~2) could cause
a companion star to have a different reddening value than the hot component.
Figure~4 compares the nebular extinction measurements (taken from CKS and
Kingsburgh \& Barlow 1994) with reddenings derived from our WFPC2 photometry
using the assumption that the intrinsic ${\rm F555W} - {\rm F814W}$ color of
the central star is $-0.40$.
For the figure, $E$(F555W$-$F814W)
has been transformed to $E(B$$-$$V)$ using the extinction table for an
O6 star given by Holtzman {\refit et~al.\ } (1995).
In the figure, M~2-9 and NGC~2022 have not been
plotted, since both of their nuclei are marginally non-stellar on our {\sl HST\/}\
frames, and are probably surrounded by thick circumstellar dust. In addition,
PNNs with known composite or late-type spectra have been omitted
from the diagram. In those cases where the PN has both a radio-flux and
Balmer-decrement value for $c$, we have used the mean of the two
measurements (except for Mz~2, whose Balmer-decrement extinction of zero is
clearly unphysical given the object's other properties).
The line in Figure~4 plots the relation between $E(B-V)$ and $c$ derived
by reddening a 100,000~K Planck curve with the Cardelli, Clayton, \& Mathis
(1989) extinction law, folding the resultant energy distribution through the
F555W and F814W system response curves, and transforming from
$E(\rm F555W-F814W)$ to $E(B-V)$ using the table of Holtzman {\refit et~al.\ } (1995).
This relation has a slight curvature, due to shifts in effective filter
wavelength with increasing extinction ({\refit e.g.,}\ Kaler \& Lutz 1985). An excellent
approximation to this curve is
\begin{equation}
E({\rm F555W-F814W}) = 0.913 \, c - 0.012 \, c^2.
\end{equation}
As can be seen in Figure~4, this law fits the data reasonably well in the
mean, as the average difference between the two reddening estimators is
$\Delta E(B$$-$$V) = 0.00 \pm 0.02$. However, it is also obvious that
there is a substantial amount of scatter; the dispersion about the relation
is $\sigma_{E(B-V)} = 0.16$~mag. Interestingly, this scatter is not confined
to highly reddened PNe: even for those objects with $c < 0.2$
the scatter is still substantial, $\sigma_{E(B-V)} = 0.13$.
Without additional information, it is impossible to determine whether
the scatter in Figure~4 is due principally to errors in the nebular
extinction values or uncertainties associated with our PNN colors. In some
individual cases, we can test whether the nebular extinction is
reasonable by computing the dereddened color of the central star: if the
application of $c$ results in a color that is significantly bluer than the
Rayleigh-Jeans limit, then the extinction has obviously been overestimated.
However, from the data at hand, there is no reason to believe that one
extinction estimate is better than the other. For simplicity, we therefore
assume that both methods have similar uncertainties, of order $\sigma_{E(V-I)}
\simeq 0.11$~mag. If a planetary nebula has both a color-estimated reddening
and a reasonable value for the nebular reddening, then we use the mean of
these two numbers, and assign an uncertainty of $\sigma_{E(V-I)} = 0.11 /
\sqrt{2} = 0.08$~mag. If no measurements of the nebula exist, if the value
of $c$ leads to an implausible color for the central star, or if the
PNN is known to have a composite or late-type spectrum (or poor photometry),
then we use a reddening based solely on the remaining valid method, and assume
$\sigma_{E(V-I)} = 0.11$~mag.
\section{Transformation to $V$ and $I$}
The final step before deriving PNN distances from binaries is to transform
the photometric systems defined by the WF/PC and WFPC2 filters to the
Johnson-Kron-Cousins $V$ and $I$ system as defined by the standard stars of
Landolt (1983, 1992). For the WFPC2 data, this was done using the equations
given by Holtzman {\refit et~al.\ } (1995), which should be good to $\sim 2\%$ for
colors in the range $-0.3 < V - I < 1.5$. Note that the bluest of our PNNs
actually lie slightly outside this range at $V$$-$$I \approx -0.4$, and the
transformations for these stars are not formally applicable. However, as
the Holtzman {\refit et~al.\ } data for the extremely blue stars of $\omega$~Cen and the
blue spectrophotometric standards Grw~$+70^\circ$5824, Feige~110, and
AGK~$+81^\circ$266 show, the equations relating F555W and F814W to $V$ and $I$
are well behaved at the blue end, and produce residuals that are no worse than
those for stars of moderate color. Hence our mild extrapolation for the
central stars should be valid.
Transformation equations are also available for the original WF/PC filter
set (cf.~Harris {\refit et~al.\ } 1991; Saha {\refit et~al.\ } 1994). However, these relations are
only defined for stars with $B$$-$$V$ colors between 0.0 and 1.6. Since many
of our nuclei are much bluer than this, we chose not to use these equations
directly. Instead, we re-derived the transformations for $V$ and $I$ by
combining the red-star data of Harris {\refit et~al.\ } (1991) with observations of
his six blue Landolt standards: G~162-66, GD~108, G~163-50, G~93-48,
Feige~67, and SA~114-750. The resulting transformation equations, which
are valid in the range $-0.5 < {\rm F555W} - {\rm F785LP} < 3$ are
\begin{equation}
{\rm F785LP} - I = 0.054 - 0.154 \ (V - I) - 0.002 \ (V - I)^2 \label{eq1}
\end{equation}
\begin{equation}
{\rm F555W} - V = 0.001 + 0.061 \ (V - I) - 0.009 \ (V - I)^2. \label{eq2}
\end{equation}
Both equations have r.m.s.~residuals of 0.023~mag.
We note here that ground-based photometry of central stars within
high-surface-brightness nebulae can be extremely difficult, due to the
bright, irregularly distributed flux from their surrounding nebulae
(see Ciardullo \& Bond 1996 for a discussion of this well-known problem).
However, the superior resolving power of the {\sl HST\/}\/ reduces this
problem enormously. Consequently, the $V$ and $I$ magnitudes
obtained in our survey represent a useful new database for future
PN analyses, even for the majority of stars that are not
resolved binaries.
Our measured PNN $V$ magnitudes and $V$$-$$I$ colors
are listed in Table~4; for reference, the nebular-based values of $c$
(which are taken from literature measurements of the Balmer-decrement and/or
radio flux density) are also tabulated. Table~5 lists the PNN $I$ magnitudes
for those objects not observed in $V$; except for the composite and late-type
stars A~46, He 1-5, and HFG~1, these data were transformed
to the standard system using the nebular-derived value of $c$ and the
assumption that the intrinsic $V-I$ color of the objects is $-0.4$.
(Abell 46, He 1-5, HFG~1 were transformed using ground-based colors.)
In most cases, the magnitudes listed in the tables are accurate to better than
$\sim 0.05$~mag, and represent mean values derived from our two exposures.
In a few cases, partially saturated stellar images compromised our ability to
derive an accurate magnitude for the PNN{}. For the objects where this
occurred,
we double-checked our measurements by deriving both
a PSF magnitude and an aperture-photometry magnitude, and intercomparing
the results computed from our long-exposure frame with those obtained from the
short exposure. Those PNNs with uncertain magnitudes and colors are
noted in Tables~4 and 5 with colons.
Two central stars, those of IC~4997 and NGC~5315,
were so badly overexposed that we were
unable to derive
$V$ and $I$ magnitudes; and two more, NGC~6309 and NGC~6369, were too badly
overexposed in the $I$ band only. For A~82 we are uncertain of the identity
of the central star, as discussed in detail below.
Figure~5 compares our {\sl HST\/}\/ $V$ magnitudes to ground-based measurements
obtained by Kaler and collaborators (Shaw \& Kaler
1985, 1989; Jacoby \& Kaler 1989) and by Tylenda {\refit et~al.\ } (1991). The figure
demonstrates that for PNNs with magnitudes brighter than $V \sim 15.4$, the
agreement is generally good. There is no systematic error between
our measurements and those of Tylenda {{\null} et al.,} as the mean difference
between the two magnitude systems is $0.00 \pm 0.06$~mag. Moreover, although
our {\sl HST\/}\/ magnitudes are systematically fainter than the Kaler {\refit et~al.\ }
measurements, the offset is small: when the highly discrepant objects IC~3568,
NGC~7662, and NGC~7008 (which we resolve as a binary with two comparably
bright components) are omitted, the mean difference between the two systems
is $0.05 \pm 0.03$~mag. We note that, in general, the scatter between our
measurements and those of Kaler {\refit et~al.\ } is significantly smaller ($\sigma \simeq
0.13$~mag) than that for Tylenda {\refit et~al.\ } ($\sigma \simeq 0.32$~mag). But,
more importantly, both sets of residuals are larger than the internal errors
of the measurements would indicate. Some of this additional scatter
undoubtedly comes from intrinsic variability in the PNNs themselves, as
Bond \& Ciardullo (1990) and others have shown that $\sim 0.1$~mag variability
in PN central stars is not uncommon.
For PNNs fainter than $V \simeq 15.4$, the agreement between the {\sl HST\/}\/ data
and the ground-based measurements is considerably poorer. This is
presumably due to the increased importance of nebular contamination
in the ground-based photometry. The effect is particularly noticeable in the
Kaler {\refit et~al.\ } data, where $V$-band fluxes derived from aperture photometry
are overestimated by as much as $\sim 3$~mag.
\section{Physical Associations and Optical Doubles}
The goal of this paper is to determine distances to PNe with resolved
binary nuclei by fitting the companion stars to the main sequence.
To do this, one should, in principle, use a main sequence appropriate for
the metallicity of each individual nebula. However, of the candidates
listed in Table~3, fewer than half have nebular abundance measurements, and
of those, only $\sim 5$ have reliable data on atomic species that can
reasonably be assumed to have been unaffected by nuclear processing.
We have therefore chosen not to attempt any metallicity correction. Moreover,
since most PNe arise from an old-disk population, we have also
chosen not to use a main sequence based on observations of young, metal-rich
clusters such as the Hyades or Pleiades. Instead for our distance estimates,
we use an $M_V$, $V$$-$$I$ main sequence that is derived from a spline-fit to
a color-magnitude diagram of old-disk field stars. This relation, which
has kindly been provided to us by H.~Harris (1998), is based on
U.S. Naval Observatory (USNO) CCD parallaxes of faint field stars
(Monet {\refit et~al.\ } 1992; Dahn 1993; plus recent unpublished measurements),
USNO photographic parallaxes of bright stars (Dahn {\refit et~al.\ } 1988),
and a combination of data on large-parallax stars from other observatories.
Since the dataset excludes known halo objects, high-velocity stars, and
binaries, the resultant main sequence should be applicable directly to our PN
companions. The adopted USNO main sequence is given in Table~6.
As mentioned above, our candidate binary PNNs were selected purely on the
basis of spatial coincidence. Hence it is likely that a few optical doubles
are mixed in with the true physical pairs. Although it is impossible to
state with complete certainty whether any individual object is, or is not,
a binary, we can make some plausibility arguments based on the inferred
properties of the stars and nebulae.
Table~7 lists inferred properties of all of our candidate binaries, grouped
according to our belief that the physical association is probable, possible,
or doubtful. The reasons for our categorization are given case-by-case below.
To derive the distance moduli given in column~3, we dereddened the stars
using the adopted $E(B$$-$$V)$ of column~2 and the prescriptions given
by Harris {\refit et~al.\ } (1991) and Holtzman {\refit et~al.\ } (1995). We then fit the dereddened
companion(s) to the USNO main sequence of Table~6. Column~4 lists the
formal errors in the derived distance moduli; these were calculated by
combining in quadrature our photometric errors, the errors associated with
the PNN reddening estimates, and the uncertainty in the main-sequence $M_V$
value. Note that this last term, which we assume to be $\sim 0.3$~mag,
usually dominates the error. Due to the spread of field-star metallicities
and the effects of stellar evolution, the observed main sequence in the solar
neighborhood has at least a spread of 0.3~mag (Perryman {\refit et~al.\ } 1995; Jaschek
\& G\'omez 1998). Consequently, although the mean main sequence at any color
may be well defined, individual distance determinations which are based only
on two-color photometry must have at least this amount of error. In the future
it should be possible to refine our distance estimates via metallicities
deduced from spectroscopy and/or Str\"omgren photometry of the companion
stars, but for the present, our individual PN distance measurements carry this
substantial uncertainty.
Columns~6 through 9 in Table~7 give various properties of the stars
and nebulae under the assumption the companion stars are, indeed,
physically associated with their PNNs. Column~6 is the projected
PNN-companion star separation in astronomical units, column~7 is the
physical size of the PNN's nebula as derived from the angular diameters listed
in Acker {\refit et~al.\ } (1992), column~8 is the absolute $V$ magnitude of the PNN,
and column~9 is the nebula's absolute [\ion{O}{3}] $\lambda 5007$ magnitude,
based on the line fluxes given by Acker {\refit et~al.\ } (1992). Following
Jacoby (1989), we define
\begin{equation}
M_{5007} = -2.5 \log F_{5007} - 13.74,
\end{equation}
where the monochromatic flux, $F_{5007}$, is given in ergs~cm$^{-2}$~s$^{-1}$.
A discussion of the individual objects appears below.
\subsection{Probable Physical Pairs}
{\it Abell 31:\/} The central star of Abell~31 lies within an extremely
large (diameter $970\arcsec$) nebula, and has a very faint, close
companion that is detected only on our $I$ frames. The extremely
red color of the companion, $(V-I)_0 \gtrsim 3.2$, places an interesting limit
on the distance: in order to be on the main sequence, the companion star must
be closer than $\sim 440$~pc. This compares to statistical distance
estimates that range from 230~pc (CKS) to 1~kpc (Zhang 1995). The small
projected separation derived for the pair ($< 115$~A.U.) and the very low
probability for chance superposition (0.07\%) argue strongly for additional
observations.
{\it Abell 33:\/} The companion to Abell~33 was first detected by
Cudworth (1973); the separation and position angle of the pair has not
changed significantly since then. The low probability of a chance
superposition and the reasonable implied parameters of the system
suggest that the stars are physically associated. Additional evidence for this
conclusion comes from the approximate agreement between our distance of
1.2~kpc and the statistical distance estimates of $\sim 0.7$~kpc (CKS),
1.6~kpc (Maciel 1984), and 2.9~kpc (Zhang 1995).
We note here that Abell~33 (and Abell~24) have the largest reddening
discrepancies in our sample. According to the observed $V$$-$$I$ color of
the central star, the extinction towards Abell~33 should be close to zero.
However, Kaler, Shaw, \& Kwitter (1990) have used the nebular emission
lines to infer a total {H$\beta$\ } extinction of almost a magnitude. Because
of the object's high galactic latitude, the low extinction values derived
by earlier emission-line studies (Chopinet \& Lortet-Zuckermann 1976; Kaler
1983), and the negative reddening implied by the ratio of radio to H$\beta$
flux (CKS), we have chosen to ignore the Balmer-line extinction measurement
for this object, and have used $E(V$$-$$I) = 0$ in our calculations.
{\it K 1-14:\/} Prints of the Palomar Sky Survey (POSS) reveal that a pair
of stars separated by $9\farcs1$ is at the center of this nebula.
H.E.B. noted this fact $\sim 20$~years ago, as did Kaler
(1981), and on this
basis we included the object in our program. However, Kaler \& Feibelman
(1985) concluded that neither of these stars is the PNN: using both {\it
International Ultraviolet Explorer\/} ({\it IUE\/}) observations and a
large-scale ground-based plate provided to them by F.~Sabbadin, these authors
concluded that the actual central star of K~1-14 is a fainter third object,
lying {\it between\/} the two POSS stars. Our {\sl HST\/}\/ images confirm that
the Kaler-Feibelman conclusion is correct: an extremely blue object lies
$\sim 2\farcs 4$ southwest of the brighter POSS star. (The PNN is invisible
on the POSS prints due to the overlapping images of the field stars.)
Remarkably, the hot star itself has a very close, even fainter companion
$0\farcs 36$ away. The very small separation of {\it this\/} pair, coupled
with the low stellar density in the field, argue strongly for a physical
association. Our main-sequence-fitting distance of $\sim 3$~kpc is in
reasonable agreement with that determined from statistical methods (3.4~kpc,
CKS; 5.3~kpc, Maciel 1984).
{\it K 1-22:\/} This system has a number of remarkable parallels to K~1-14.
The discoverer of the PN (Kohoutek 1971) noted that three stars appear near
the center of the nebula, and proposed that the brightest of these was the
PNN{}. Smith \& Gull (1975) confirmed that this star is very blue, but
Kaler \& Feibelman (1985) noted that the visual luminosity of the star
was too large to match the flux distribution extrapolated from the
{\sl IUE\/}\/ measurements. Instead Kaler \& Feibelman suggested that a fainter
object $4''$ east of Kohoutek's candidate was the PN's true central star.
Our {\sl HST\/}\/ frames confirm that the original Kohoutek-Smith-Gull star is indeed
the PNN, but its visual flux is augmented by a very close visual companion
$0\farcs 35$ away. In fact, at $V=17.13$, the companion is so red
($V-I=1.12$) that it is actually brighter than the PNN in the $I$ band.
Given the low stellar density in the field, the pair almost certainly form
a bound system. Our derived distance of 1.3~kpc is reasonably
consistent with distances based on statistical techniques (1.0~kpc, CKS;
3.4~kpc, Zhang 1995).
{\it K 1-27:\/} The faint companion to K~1-27 has the color of a late A~star;
consequently if the star is on the main sequence, its faint apparent magnitude
($V \simeq 21.3$) implies an implausibly large distance (55~kpc). There is,
however, another possible solution for this system: the companion star could
be a white dwarf. Instead of placing the secondary on the main sequence, we
obtained the distance given in Table~7 by putting the companion on the
white-dwarf cooling sequence defined by the hydrogen-rich white-dwarf models
of Bergeron, Wesemael, \& Beauchamp (1995). The locus of points defined by
these models is in excellent agreement with that observed for field white
dwarfs (Monet {\refit et~al.\ } 1992), and, by adopting the curve, we derive a white-dwarf
absolute magnitude of $M_V = 12.77$. If we use this value, and ignore the
nebular Balmer-line extinction estimate $c = 0.28$ (which leads to an
unphysically blue color for the central star), then we derive a distance to
the system of $\sim 470$~pc.
There is only one other distance estimate, statistical or otherwise, for
K~1-27. By modeling the absorption lines from the hydrogen-deficient central
star, Rauch, K\"oppen, \& Werner (1994) derived a distance to the object of
$1.29^{+1.05}_{-0.58}$~kpc. This number, however, assumes $c = 0.28$; with
a more reasonable value of $c = 0.08$, their distance decreases by a factor
of $\sim 1.2$, and comes into marginal agreement with ours. In most
other respects, our white-dwarf hypothesis is reasonable. At an
absolute magnitude of $M_V = 12.77$, the cooling age of a $0.575 M_\odot$
companion white dwarf is $\sim 1$~Gyr (cf.~Bergeron, Wesemael, \&
Beauchamp 1995), and thus the star would not be in a particularly rapid
phase of evolution. Similarly, if the distance is indeed 0.47~kpc, then
the derived values for the nebular size (0.1~pc) and binary
separation (260~A.U.) are also plausible. The only potential problem lies in
the extremely small amount of flux radiated by the nebula in [\ion{O}{3}]
$\lambda 5007$. At an absolute [\ion{O}{3}] $\lambda 5007$ magnitude of $M_{5007}
\sim 8.7$, the nebula would be a full 13.1~mag fainter than the bright end of
the [\ion{O}{3}] $\lambda 5007$ planetary nebula luminosity function (Ciardullo
{\refit et~al.\ } 1989; Jacoby {\refit et~al.\ } 1992), and the intrinsically faintest [\ion{O}{3}]
$\lambda 5007$ source in the Strasbourg-ESO planetary nebula catalog (Acker
1992). Note, however, that the nebular and central-star properties of K~1-27
(Henize \& Fairall 1981; M\'endez, Kudritzki, \& Simon 1985) are very similar
to those of Abell~36, an extremely high-excitation object which, from its
[\ion{O}{3}] line flux and statistical distance, {\it is\/} the faintest [\ion{O}{3}]
$\lambda 5007$ source currently in the catalog ($M_{5007} \sim 7.2$). In
fact, both Henize \& Fairall and M\'endez {\refit et~al.\ } have remarked that the
properties of the K~1-27 nebula and central star are more extreme than those
of Abell~36, and a large fraction of the nebula's oxygen is in O~IV{}.
Consequently, the intrinsically small amount of [\ion{O}{3}] $\lambda 5007$
emission is entirely reasonable. We consider this a likely white dwarf-PNN
binary system.
{\it Mz~2:\/} This PN is 3 degrees from the Galactic plane, and only 31 degrees
from the Galactic center. Thus, the field-star density in the region
is extremely high: there are over 3400 stars recorded on our Wide Field
and Planetary Camera frames. Nevertheless, only five of these stars are as
bright as, or brighter than, our putative companion. The relative scarcity of
$I = 15.8$ stars, and the small ($0\farcs 28$) separation between the PNN and
the companion, leads to a very high probability (99.99\%) of a physical
association.
Our distance to Mz~2 of $\sim 2.2$~kpc is in good agreement with most
statistical distance estimates (2.3~kpc, CKS; 2.4~kpc, Van de Steene \&
Zijlstra 1994; 2.7~kpc, Zhang 1995). Kingsburgh \& English (1992) derive a
somewhat larger distance ($\sim 5$~kpc) from the PN's location on the
[\ion{O}{2}] density-ionized mass relation, but the subsequent classification
of the object as a Type~I PN (Perinotto {\refit et~al.\ } 1994) vitiates this analysis.
{\it NGC 1535:\/} The companion to this well-studied PNN has the colors
of an early G~star. If the pair form a bound system, then the projected
separation between the two stars is $\sim 2400$~A.U., and the distance
to the binary is $\sim 2.3$~kpc. This value is in excellent agreement not
only with most modern statistical distances (2.3~kpc, CKS; 2.0~kpc, Van de
Steene \& Zijlstra 1994; 2.1~kpc, Zhang 1995), but with the distance derived
from the non-LTE model-atmosphere analysis of its central star (2.0~kpc,
M\'endez, Kudritzki, \& Herrero 1992). This consistency, along with
the small probability of a chance superposition, supports the interpretation
that this is a true physical system.
{\it NGC 3132:\/} The binary nature of this PNN was first discovered by
Kohoutek \& Laustsen (1977). The companion star has an A0~V spectral type
(Lutz 1977), which implies that the PN was ejected from an even more massive
progenitor. The hot star, which is only marginally resolved from the
ground, is easily measured on our PC frame, and is 5.65~mag fainter in $V$
than its companion; if we place the A~star on the main sequence, we
obtain a distance to the system of $\sim 0.8$~kpc. This is consistent with
estimates based on the interstellar reddening along the line of sight
(Gathier, Pottasch, \& Pel 1986) and the ground-based spectroscopic parallax
of the A~star (Pottasch 1980).
Unfortunately, there are two caveats that accompany our distance
measurement. The first arises out of the definition of our field-star
main sequence. For a PN to exist, the age of the system must be at least
$\sim 5 \times 10^7$~years (Bressan {\refit et~al.\ } 1993); this is a non-negligible
fraction of the main-sequence lifetime of the A~star. Consequently, some
main-sequence evolution must have occurred, and the companion could be up
to $\sim 0.2$~mag brighter than its zero-age main sequence luminosity. While
the effect is partially mitigated by our use of a field-star main sequence
rather than a zero-age main sequence, it is still likely that the companion
is older than a typical A-type field star. Fortunately, the effect is small:
the difference between the expected mean magnitude of a group of A stars
of all ages, and that of a sample of stars with ages of at least
$\sim 5 \times 10^7$~years is only $\sim 0.05$~mag.
The second problem comes from the uncertainty in our reddening estimation.
On the lower main sequence, the $V$ vs.~$V$$-$$I$ reddening vector is roughly
parallel to the main sequence. Consequently a small error in the extinction
makes little difference to a distance derivation: both $V_0$ and $M_V$
are modified in a similar fashion. For A~stars, however, this is not the
case, as the slope of the main sequence is significantly steeper. The
result is that a small uncertainty in reddening translates into a large
uncertainty in distance. In the case of NGC~3132, the uncertainty in our
color-based reddening estimate, $\sigma_{E(V-I)} \sim 0.11$~mag, translates
into a $\sim 0.64$~mag uncertainty in distance modulus. This dominates
the error given in Table~7. Str\"omgren photometry of the A~star would
improve the distance estimate considerably.
{\it NGC 7008:\/} It is somewhat surprising that the composite nature of
the nucleus of NGC~7008 was not detected prior to our survey. Although the
angular separation between the PNN and the companion is small ($0\farcs 4$),
the magnitude difference between the two stars is only $\sim 0.5$~mag in
$V$, and in $I$ the light from the companion star actually dominates. By
placing the companion on the main sequence, we obtain a distance of
$\sim 0.4$~kpc, and an implied stellar separation of $\sim 160$~A.U.{}
This compares to a distance of $\sim 1.1$~kpc estimated from extinction
measurements along the line of sight (Pottasch 1983), and the CKS statistical
distance of $\sim 0.9$~kpc. Because the probability of a chance superposition
of a $V \sim 14$~mag star near the PNN is extremely small, we consider this
a good binary-star candidate.
There is one caveat to our distance, however. On both of our F814W frames,
the point-spread function for NGC 7008's companion star appears slightly
broader than expected from a normal single star. The effect is not large and
we cannot rule out the existence of an instrumental problem. However, it is
possible that the companion is itself a binary, making the PNN a hierarchical
triple. If this is the case, our distance to the object is an overestimate.
{\it Sp 3:\/} The companion in this system has the color of an F-type star
and is only $0\farcs 3$ from the PNN; when we place it on the main
sequence, we derive a distance of $\sim 2.4$~kpc, in good agreement with the
CKS statistical distance of 1.9~kpc. This consistency, along with the small
probability of superposition, indicates that this is a true physical
association.
\subsection{Possible Physical Pairs}
{\it Abell 7:\/} Our 200-s $I$-band Planetary Camera image shows only two
stars on the frame: the PNN and a very faint companion only $0\farcs 9$
away. Since the stellar density in this field is low (it is at
$b = -30^\circ$), there is a very high probability that
the stars are physically associated. Unfortunately, the companion is
$\sim 5.3$~mag fainter than the central star in $I$ and so red that
it was not detected on our $V$~frames. This implies a $V$$-$$I$
color redder than 1.21 and an absolute magnitude fainter than $M_V = 6.7$.
From main-sequence fitting, we can derive only an upper limit to
Abell~7's distance of $\sim 13$~kpc. This limit is not
particularly useful, as the size of the nebula ($760\arcsec$) and the
central star's white-dwarf spectrum (Liebert 1980) demand that the
object be quite nearby. If the CKS statistical distance of $\sim 200$~pc is
accurate, then the projected separation between the PNN and the companion
star is only $\sim 200$~A.U{}. Deeper images are desirable to investigate
this interesting object further. In the meantime we can only classify A~7 as a
possible physical binary.
{\it Abell 30:\/} The companion to Abell~30 was found from the ground by
Cudworth (1973). The position angle has not changed significantly over the
ensuing $\sim 25$~years, and our main-sequence fitting distance ($\sim 2$~kpc)
is consistent with the CKS statistical distance of 1.7~kpc. On the
other hand, the relatively large separation ($5\farcs 25$) does allow
a $\sim2\%$ probability of a chance superposition. Moreover, at a distance
of 2~kpc, the derived physical separation of the pair ($\sim 10,600$~A.U.)
is starting to approach the observed 0.1~pc cutoff imposed by the Galactic
tidal field (cf.~Bahcall \& Soneira 1981; Latham {\refit et~al.\ } 1984). We therefore
categorize A~30 only as a ``possible'' physical association.
{\it Abell~63:\/} The central star of Abell~63 is the 11-hour eclipsing
binary UU~Sge (Bond, Liller, \& Mannery 1978); the PNN's nearby companion
($2\farcs8$ away) was first noted during photoelectric observations by
Krzeminski (1976). In spite of the PN's low galactic latitude, $b=-3^\circ$,
we find only a $1.5\%$ probability that the resolved companion is a
chance superposition. Based on the nebular Balmer decrement, the extinction
to the object is $c=0.71 \pm 0.10$ (Walton, Walsh, \& Pottasch 1993); this
number is in excellent agreement with the value obtained by modeling the
2200~\AA\ interstellar
absorption feature (Walton {\refit et~al.\ } 1993; Pollacco \&
Bell 1993). When we combine this extinction estimate with our {\sl HST\/}\
photometry, we derive a distance to the companion star of $\sim 1.2$~kpc.
From the run of stellar reddening versus distance in that part of the sky,
a foreground extinction of $c \sim 0.7$ ($E(B$$-$$V) = 0.5$) implies a
distance of $\sim 1$~kpc (cf.~Bond {\refit et~al.\ } 1978), in good agreement with
our findings. However, analyses of the spectral type and color of the
back hemisphere of the extremely close eclipsing companion yield values that
are two to three times larger ($\sim 3.6$~kpc, Walton {\refit et~al.\ }
1993; $3.2 \pm 0.6$~kpc, Pollacco \& Bell 1993; $2.4 \pm 0.4$~kpc,
Bell, Pollacco, \& Hilditch 1994). These distances could be
overestimates, if some of the extreme heating effects on the companion
``leak'' around to the unilluminated back side. Unfortunately, without
improved observations, we can only identify this as a possible physical
system.
{\it IC 4637:\/} This new pair---which remarkably was never noted from the
ground---was discovered on our original Cycle~3 F785LP WF/PC frame and
followed up with WFPC2 (F555W and F814W) observations. The implied
separation between the PNN and its companion, $\sim 1200$~A.U., is large,
but not excluded by any means; similarly, the derived nebular size of
$\sim 0.05$~pc is small, but again not unreasonable. Interestingly, the
statistical distances to this object vary widely, from 0.8~kpc (Amnuel
{\refit et~al.\ } 1984) to 2.3~kpc (CKS), but all estimates lie on the high side of
our 0.5~kpc value. Furthermore, the object's only individual distance
determination, the non-LTE model-atmosphere value of M\'endez, Kudritzki,
\& Herrero (1992), is larger still (3.3~kpc). We are therefore led to
classify this object only as a possible physical binary.
{\it NGC 2392:\/} Like Abell 7, NGC~2392 has a faint companion which is
barely detected on our $I$~frames, and is invisible in $V$. Our
upper limit to the distance, $\sim 6.4$~kpc, is not particularly useful,
since both non-LTE model atmospheres (M\'endez {\refit et~al.\ } 1992)
and statistical techniques (CKS; Van de Steene \& Zijlstra 1994; Zhang 1995)
place the object closer than $\sim 2$~kpc. Deeper imaging could provide a
photometric distance, but for the present, the lack of a $V$ magnitude
forces us to categorize the object as a ``possible'' physical association.
{\it NGC 2610:\/} Again, there is a close ($0\farcs 6$), extremely faint
companion to this object that could not be detected on our $V$ images.
A much deeper image is needed to produce a meaningful distance estimate.
We again categorize this as a ``possible'' physical system.
\subsection{Doubtful Physical Pairs}
{\it Abell 24:\/} The visual companion to this star was first noted
by Cudworth (1973). We measure a separation of $3\farcs33$, in good
agreement with the $3\farcs4$ reported by Cudworth.
However, our measurement for the position angle of the binary
differs by $5^\circ$ from his value; this
offset marginally exceeds the uncertainty in the old measurement
(Cudworth 1997). Since for any plausible distance, this change in position
angle is larger than is possible from binary orbital motion, the difference
argues against the existence of a physical association. Other facts
suggestive of a chance superposition include the extremely
large diameter derived for the nebula ($\sim 4$~pc) and the factor of
five difference between our putative distance ($\sim 2.4$~kpc) and distances
derived from statistical techniques ({\refit e.g.,}\ 0.52~kpc, CKS). Despite the
fact that the formal probability of a chance superposition is only $\sim 2\%$,
we believe that this object is an optical double.
{\it NGC 650-1:\/} The companion star of this object was first noticed by
Cudworth (1973). However, our {\sl HST\/}\ images reveal that the companion
is itself a very close double, with a separation of only $0\farcs 16$.
Given the field star density of the region, the two stars of the companion
almost certainly form a physical system, as the probability of a $0\farcs 16$
chance superposition is less than 0.1\%. However, associating the
companion binary with the PNN is more problematical.
The separation and position angle of the companion pair relative to the PNN
is the same as it was 25 years ago; thus, despite the fact that the
probability of a chance association is $\sim5\%$,
this may argue for a physical
association. However, if the system is a hierarchical triple, then it is a
very peculiar one: although both components of the companion have the same
color, $(V-I)_0 \approx 0.87$, one is 0.8~mag brighter than the other. In
other words, both stars cannot be on the main sequence. If the PNN is
associated with this pair then two of the three stars of this non-interacting
trinary happen to be in a phase of rapid evolution.
The hypothesis of a bound triple system runs into further problems when one
considers the age and metallicity of the stars. NGC~650-1 is a metal-rich
Type~I planetary nebula, with a most likely progenitor mass $M \gtrsim 2
M_\odot$ (Peimbert \& Torres-Peimbert 1983). Yet if we assume a solar-like
metallicity and place the fainter component of the close pair on the main
sequence, then the position of the brighter component in the color-magnitude
diagram demands that the system be impossibly old. As the isochrones in
Figure~6 show, both of the close components could lie on or close to the same
isochrone if the $V-I$ colors were near the blue ends of their error bars, but
the age of the pair would still be far in excess of that of the PNN{}. A
lower age for the close pair would be possible if a substantial amount of
additional reddening were adopted. However, if that were the case, then
the reddening would greatly exceed that of the PNN itself, and we would be
forced to conclude that the two systems were unrelated. Another alternative
is that the fainter component of the close pair is on the main sequence, and
the brighter component is itself an unresolved binary containing two
equal-brightness main-sequence stars; although not impossible, this option
would imply the unlikely conclusion that the close pair is actually a
hierarchical triple containing three stars of essentially identical mass.
The distance of 4.1~kpc given in Table~7 was derived by fitting the fainter
component of the close pair to the main sequence; it exceeds the statistical
distances (CKS: 0.7~kpc; Zhang 1995: 1.6~kpc; Van de Steene \& Zijlstra 1994:
1.3~kpc) by a substantial amount. We therefore suspect that the very close
pair, whatever its astrophysical explanation, is probably significantly more
distant than---and older than---the PNN, and thus not physically
associated.
{\it PuWe 1:\/} A slightly brighter companion to the central star of this
object is visible on the Palomar Sky Survey at a separation of $5''$
(Purgathofer \& Weinberger 1980), but our WF/PC observations reveal that
the companion is itself resolved into two stars with a separation of
$0\farcs 6$ arcsec. Unlike the NGC~650-1 system, both cool components
are apparently on the main sequence, as they produce a consistent set of
distances. Since the PN is an extremely large object that extends over
$10'$ on the sky (Purgathofer \& Weinberger 1980), we expect it to be
relatively nearby, and our derived distance of $\sim 280$~pc does not
contradict this hypothesis.
Nevertheless, the PuWe~1 PNN is probably not associated with the close pair.
Based on the stellar density in the region, and the rather large
($5\arcsec$) separation between the PNN and the companions, the likelihood of
a chance superposition in the region is relatively large, $\sim 3\%$.
Moreover, PuWe~1 is one of the few PNNs with a reliable trigonometric parallax
measurement, $2.3 \pm 0.4$~mas (Harris {\refit et~al.\ } 1997). This is more than $2
\sigma$ smaller than we would predict based on our companion star
measurements. Even more telling is that the proper motion of the close pair
appears to be significantly different from that of the PNN (Harris 1998).
Thus, we believe the stars are not physically associated.
\subsection{Noteworthy Optical Doubles and Unresolved Objects}
Several other stars in our survey were included because they were identified
as candidate visual binaries on the basis of ground-based observations, but we
have concluded that they are not physical pairs. In addition, a few other
objects are definitely binaries, based on their composite spectra and/or red
colors, but we did not resolve them. These objects are discussed in this
subsection.
{\it Abell~78:\/}
In his survey of the nebular features of this PN, Jacoby (1979) noted that
an extended patch of H$\alpha$ emission $\sim 10\farcs 2$ from the
PNN was spatially coincident with a star, but that there was no other
evidence for a physical association. Our star counts in the region
suggest that the ``companion'' is most likely an unrelated field star which
is merely projected onto the nebula.
{\it Abell 82:\/} A~82 has a diameter of about $90''$, and was included in our
program as a candidate binary because, as recounted by Kaler \& Feibelman
(1985), there is a relatively bright late-type star at the center of the
nebula. Kaler \& Feibelman's {\sl IUE\/}\/ spectra did reveal a weak, apparently
reddened, UV continuum within the $10''\times 20''$ aperture of the
instrument, but the derived color temperature was much too low to explain the
nebula's ionization. The central object has a K-type spectrum (Kaler \&
Feibelman 1985) and colors of $V=14.90$, $B$$-$$V=1.28$, $V$$-$$I=1.36$
(Kwitter, Jacoby, \& Lydon 1988, and our {\sl HST\/}\/ measurements). Kwitter {\refit et~al.\ }
have suggested that the image of the K~star masks the true PNN on ground-based
photographs (cf.~the case of K~1-14 above), but we find no evidence on our
{\sl HST\/}\/ frames for any blue object near the center of the nebula.
Alternatively, Kaler \& Feibelman (1985) have proposed that a faint object
$6''$ northwest of the K~star is the PNN, but our photometry ($V=18.15$,
$V-I=1.10$) demonstrates that this star is not blue either. In fact, the
bluest object on our PC frame is a $V=12.85$, $V$$-$$I=0.35$ star located
$18''$ southeast of the K~star. This star might have been within the {\sl IUE\/}\/
aperture if the pointing was slightly inaccurate, and thus could be the
reddened source detected by Kaler \& Feibelman. However, it is too far
off-center to be the planetary's central star.
In view of the above ambiguities, we investigated the {\sl IUE\/}\/ observations in
further detail, with the aid of copies of the original observing scripts which
were kindly provided by W. A. Feibelman. Two short-wavelength spectra were
obtained, SWP~19771 (1983 Apr 20, 25~min), which shows no convincing
detection, and SWP~19908 (1983 May 5, 120~min), which detected the reddened
continuum described above, and shows that the source was well-centered in the
aperture. For both observations, a blind offset was done from a nearby
11th-mag star onto the coordinates of the center of the PN{}. Our
measurements on the Digitized Sky Survey show that the resulting pointing
would have been about $10''$ directly west of the 14th-mag K-type star, and
thus would either have missed it, or at best have had it at the edge of the
aperture. However, handwritten notes from the telescope operator indicate
that a possible additional telescope movement may have been performed onto a
``second central star,'' which we speculate could have been the 12.8-mag star
to the southeast of the K~star. In this case, the weak UV spectrum would be
nicely explained. If so, however, it appears that {\sl IUE\/}\/ never actually
observed the star at the center of the PN{}.
We are thus left with two plausible conclusions: either the true hot central
star is a faint, unresolved companion of the K-type star; or a ``born-again''
scenario (see next paragraph) for the K~star itself may have to be invoked.
Further ground- and space-based spectroscopic observations of this PNN are
urged.
{\it He~1-5\/ \rm and \it H~3-75:\/} The central star of He~1-5 is the
well-known object FG~Sge, which appears to be a PNN that has undergone a
late helium thermal pulse and has become a ``born-again'' red giant. In
agreement with spectroscopic observers ({\refit e.g.,}\ Feibelman \& Bruhweiler 1990),
and in agreement with the born-again scenario, we find no evidence for a hot
companion in our {\sl HST\/}\/ images. H~3-75 is an interesting and possibly related
object: according to Sanduleak (1984) the central star has a K-type spectrum.
{\it IUE\/} ultraviolet spectra (Bond 1993) show no trace of a hot
star, and our {\sl HST\/}\/ frames likewise show no resolved companion. The
PNN is therefore a strong candidate for another born-again giant.
Spectroscopic observations are needed.
{\it He 2-36:\/} The optical central star of He 2-36 has a spectral type of
A2~III (M\'endez 1978), and {\it IUE\/} observations suggest the presence of a
hot companion (Feibelman 1985). Our frames do not resolve the binary.
{\it M 1-2:\/} This object has a G2~Ib spectral type (O'Dell
1966), and strong forbidden and permitted emission lines reminiscent of those
seen in a planetary nebula (see Grauer \& Bond 1981 and references therein).
Our {\sl HST\/}\/ observations do not resolve any binary companion, nor do they show
a resolved nebular component. The system is probably a symbiotic-like binary
which is too compact to be resolved by {\sl HST\/}, although Feibelman (1983) has
argued from {\it IUE\/} observations that the star is surrounded by a young
planetary nebula.
{\it NGC 1514:\/} The central star is a well-known composite system,
containing a hot sdO star and an A-type companion (Kohoutek \& Hekela 1967;
Greenstein 1972). Greenstein's radial velocity measurements indicate that
the period of the system must be quite long (or perhaps that the binary is
seen nearly pole-on). Nevertheless, our {\sl HST\/}\/ images fail to resolve the
system, setting an upper limit of approximately 40 A.U. for the projected
separation.
{\it NGC 6853:\/} The binarity of the PNN was first suggested by Cudworth
(1973), who identified a $V \simeq 17$ companion $6\farcs 5$ from the central
star. Astrometric measurements (Cudworth 1977) support this contention, as
the proper motion of the companion is similar to that of the central star.
Unfortunately, given the density of field stars in the region, our probability
calculations cannot confirm this claim of binarity, as there is almost a 90\%
chance that a random field star will be projected within $6\farcs 5$ of the
PNN{}. In fact, based on the star counts, the best candidate for association
with NGC~6853 is a $V \approx 18.7$ star $1\farcs 1$ from the central star,
but even this object has a 12\% probability of being a chance superposition.
We do note that if we assume Cudworth's star is associated with the PNN, then
our {\sl HST\/}\/ photometric values of $V = 16.91$, $V$$-$$I = 1.83$, coupled with
the assumption of no foreground reddening, leads to a distance of
$430 \pm 62$~pc. This is in agreement with the distance of $380 \pm 64$~pc
recently obtain by Harris {\refit et~al.\ } (1997) from USNO parallax measurements. In
keeping with the precepts of this paper, we will not discuss this object any
further, but we urge radial-velocity measurements for the system.
{\it Th 2-A:\/} As noted in \S 2, this object was included in our program
because of a nearby companion noted on ground-based CCD frames. However,
there are over 2300 field stars present on our {\sl HST\/}\ frames, and the star
in question has a $\sim 50\%$ probability of being a chance superposition.
We therefore cannot classify it as a possible visual binary.
{\it A~34 and A~66:\/} Like Th 2-A, nearby companions were noted during
ground-based observations. The companion of A~34 has a 28\%
probability of being there by chance, and for A~66 the probability is 21\%.
{\it A~46, A~65, HFG~1, K~1-2, LoTr~5, \rm and \it NGC~2346:\/} All of these
central stars (along with A~63, discussed above) are known to be extremely
close binaries, based on their photometric variability (cf.~Bond \& Livio
1990). Not unexpectedly, none of them were resolved in our survey, and we do
not find any nearby resolved companions that have a high probability of being
physically associated.
\section{Comparison with Statistical Distance Scales}
Due to their complexity and vast range in size, luminosity, mass, and
excitation, there is no reliable method for obtaining distances to large
samples of individual Galactic planetary nebulae. As a result, in order to
investigate planetary nebulae as a class, it is necessary to rely on
statistical distance estimators. The principle behind these statistical
distances is straightforward: the emitted Balmer-line flux from an ionized
plasma depends almost exclusively on the total mass of the emitting region and
the plasma density. Consequently, if the ionized mass of a nebula can be
estimated, then its observed flux and angular size can be used to calculate
its distance. The key, of course, is to know the amount of ionized mass
contained in the nebula.
There are several prescriptions in the literature for estimating this
mass, starting with the original assumption by Shklovsky (1956) that the
ionized masses of all PNe are the same, and that all PNe are optically thin.
Other formulations include adopting an ionized mass that is (a)~linearly
proportional to nebular radius (Maciel \& Pottasch 1980), (b)~proportional to
a power of the radius (Zhang 1995), (c)~proportional to the radio brightness
temperature of the nebula (Van de Steene \& Zijlstra 1995; hereafter VdSZ),
(d)~dependent on the [\ion{O}{2}]-derived nebular density (Kingsburgh \&
Barlow 1992; Kingsburgh \& English 1992), or (e)~constant for optically thin
nebulae, but proportional to an optical-thickness parameter for denser objects
(Daub 1982; CKS){}. Once calibrated, each of these relations is capable of
producing distance estimates to large numbers of objects.
Unfortunately, the number of PNe with independently known distances, which can
therefore be used as zero-point calibrators for these methods, is extremely
small. Moreover, some of the calibrators
have distances that are themselves controversial. For example, CKS
used 19~PNe with ``well-determined'' distances to calibrate their distance
scale, while VdSZ used 23 calibrators. A comparison of the two samples,
however, reveals that only 16~PNe are common to both datasets, and of those,
two objects have adopted distances that differ by more than a factor of two!
A major reason for this dichotomy is that many of the ``well-determined'' PN
distances are based on such methods as reddening of field stars projected
near the PN line of sight (Gathier, Pottasch, \& Pel 1986), Galactic
\ion{H}{1}
absorption measurements (Gathier, Pottasch, \& Goss 1986), nebular expansion
parallaxes ({\refit e.g.,}\ Hajian {\refit et~al.\ } 1995), and non-LTE atmospheric
modeling of PN central stars ({\refit e.g.,}\ M\'endez {\refit et~al.\ } 1992). None
of these methods is unassailable, and each carries its own (possibly
substantial) uncertainty.
Our new
sample of PNe with visual-binary nuclei significantly increases the number
of objects with reliable distance measurements, and constitutes a new
and important set of data with which to calibrate PN statistical distances.
In addition to having quantifiable errors, the PNe in our sample have
distinctly different selection criteria from those measured by other methods.
PNe with interstellar-medium-based distances are mostly distant objects
in the plane of the Milky Way. Similarly, PNe with nebular-expansion
distances are objects that are bright and optically thick, while those
analyzed with non-LTE model atmospheres have highly evolved central
stars. Our wide binary stars, however, are primarily nearby objects and
objects at high galactic latitude. Consequently, our sample not only enlarges
the PN calibrator database, but also reduces the possibility of a systematic
error due to selection biases.
The usefulness of our dataset for testing statistical distance techniques is
demonstrated in Figure~7, which compares directly measured PN distances with
those from four different statistical methods. For the directly measured
distances, we use the 10 ``probable'' associations listed in Table~7 (one of
which, A~31, is only an upper limit), along with the three additional
distances to the ``possible'' associations which are not upper limits. To
these we add the new ground-based distance to NGC~246 (based on photometry of
its wide binary companion; Bond \& Ciardullo 1999), and trigonometric
distances to seven PNe derived from recent $> 3 \sigma$ parallaxes measured by
the {\it Hipparcos\/} satellite (ESA 1997) and the U.S. Naval Observatory
(Harris {\refit et~al.\ } 1997).
These accurate photometric and geometrical distances are plotted against
statistical distances computed from the 5~GHz flux measurements and
angular diameters given by Zhang \& Kwok (1993) and CKS, using the
prescriptions of CKS, VdSZ, Maciel \& Pottasch (1980), and Zhang
(1995). For reference, the data of Figure~7 are given in Table~8.
It is immediately obvious that distances from all of the statistical methods
have considerable dispersion. Compared to the binary and astrometric
distances, the CKS and Zhang estimates scatter by $\sigma \sim 1.7$~mag and
$\sim 1.8$~mag in distance modulus, respectively. The VdSZ distances exhibit
the smallest dispersion, $\sim 1.6$~mag, while the Maciel estimates have the
largest, $\sim 2.4$~mag. (This last result is not very surprising, since some
of the PNe considered here have radii outside the formal limits of the Maciel
calibration.) Interestingly, a significant amount of dispersion is
attributable to one object, PHL~932, which has a {\it Hipparcos\/} parallax
distance of $110^{+48}_{-26}$~pc, but statistical distances that range from
800~pc (CKS) to 5.0~kpc (Zhang). The central star of PHL~932 is an
exceptionally unusual object. With a sdB spectral type (M\'endez {\refit et~al.\ } 1988),
it is one of only two known PNNs of this class, lying well off the normal
post-AGB evolutionary tracks. It has been suggested that the star may have
evolved through a common-envelope binary interaction (Iben \& Tutukov 1993),
and thus the ionized mass could differ substantially from that of normal PNe.
If PHL~932 is arbitrarily disregarded, then the dispersion in the VdSZ and
Zhang errors drops dramatically to $\sim 1.1$~mag and $\sim 1.3$~mag,
respectively. Even this, however, is much larger than the errors expected
from the techniques.
Even more surprising are the zero-point offsets in the scales exhibited in the
figure. All four of the statistical methods examined here systematically
overestimate the distances to the objects in our sample. CKS come closest to
reproducing our distance scale, with estimates that are, on average, only
$\sim 25\%$ larger than the photometric and geometrical measurements. Since
this is a $1 \, \sigma$ result, their analysis is still consistent with our
measurements. The distance scales defined by VdSZ, Maciel, and Zhang,
however, are all significantly too long, with mean distance moduli that are
1.6, 1.2, and 2.7~mag larger than our own. The sizes of these offsets are
extraordinary, especially when one considers that the VdSZ and Zhang relations
work well for samples of PNe in the Galactic bulge.
Another way of looking at the problem is to compare the properties of our PNe
with those of other PNe with ``well-determined'' distances. Following VdSZ,
we plot in Figure~8 the distance-independent PN radio brightness temperature
(which is related to the CKS optical depth parameter), against the
distance-dependent quantity of PN radius. The filled circles represent
planetary nebulae with resolved binary companions, the crosses show PNe with
trigonometric-parallax distances, and the open circles are the VdSZ sample of
PNe with what they considered to be ``well-determined'' distances. It is clear
from the figure that the PNe in our new sample do not obey the rather tight
relation defined by the VdSZ calibrators; at a given radius, the binary and
astrometric PNe are systematically fainter in the radio and have a larger
amount of scatter (although the scatter is dominated by a few outliers).
We interpret Figure~8 as revealing a classical selection effect. The distances
for the PNe in the VdSZ sample come almost exclusively from reddening and
\ion{H}{1} absorption distance determinations; hence these objects are nearly
all relatively bright, high-surface-brightness PNe that can be seen at great
distances along the plane of the Milky Way. Consequently, a calibration using
this sample of objects nicely recovers the distances to Galactic bulge PNe.
The planetaries in our sample, however, are primarily nearby, faint, and
optically thin. They represent a population of objects that has apparently
received little weight in the calibration of statistical distances.
Figure 8 also points out the probable cause of a long-standing controversy
about the Galactic PN distance scale. For years, there have been two distance
scales for Milky Way planetaries. The traditional ``short'' distance scale
adopted by Cahn \& Kaler (1971) is supported by the extinction-distance
relation of Pottasch (1984), the Magellanic Cloud observations of CKS and
Webster (1969), and the Galactic bulge measurements of Stasi\'nska {\refit et~al.\ }
(1991). All of these methods use bright PNe, similar to those analyzed in the
VdSZ study. The opposing ``long'' distance scale, which is larger by a factor
of $\sim 1.5$, is supported by statistical-parallax measurements (Cudworth
1974), stellar-atmosphere models (M\'endez {\refit et~al.\ } 1992), [\ion{O}{2}]
line-ratio density estimates (Kingsburgh \& Barlow 1992; Kingsburgh \& English
1992), and number counts of PNe in other galaxies (Peimbert 1990). These
techniques study a different sample of PNe, and include local,
lower-surface-brightness objects. Based on the results of Figures~7 and 8, it
is therefore not surprising that a different distance scale is derived.
It is unfortunately less obvious---and beyond the scope of this paper---how
one could devise a new ``grand-unification'' calibration that simultaneously
handles both the lower-surface-brightness objects that prevail among the
nearby nebulae, and the brighter PNe that dominate samples like those in the
Galactic bulge and extragalactic systems. We leave this daunting task to
future workers.
\section{Conclusion}
We have successfully used a large-scale {\sl HST\/}\/ snapshot survey to find 19
resolved companions of central stars in planetary nebulae. We consider ten of
these systems to be probable physical associations, another six to be possible
associations, and the remaining three to be doubtful.
By fitting the companions to the main sequence (or in one case the white-dwarf
cooling sequence), we have derived reliable distances to the PNe. Comparison
with various statistical distance estimates reveals that all of the current
statistical methods {\it overestimate\/} the distances to our sample. A more
detailed examination suggests that the well-studied nebulae used as
calibrators for statistical methods are biased toward high-surface-brightness,
low-Galactic-latitude objects. Our sample, on the other hand, contains more
objects of lower surface brightness, which may have systematically lower
nebular masses. It will be a challenge to future refinements of the
statistical methods to include an additional correction for this effect.
The primary source of error in our PN distance measurements from resolved
binaries is the unknown metallicities of the companion stars. This
uncertainty propagates directly into the definition of the $M_V$, $V$$-$$I$
main sequence used to obtain the absolute magnitudes (and distance moduli) of
the stars. Our distances could therefore be improved substantially via
abundance analyses of the nebulae and/or stellar atmosphere analyses (or
intermediate-band photometry) of the companion stars. In addition, since
several candidate companion stars were detected only in $I$, deeper multicolor
imaging with {\sl HST\/}\/ (possibly with the restored NICMOS camera) could add
significantly to the list of PNe with direct distance determinations. Finally,
in order to confirm our candidate companion stars and identify additional
ones, proper-motion and radial-velocity measurements are needed.
\acknowledgments
We thank H.~C.~Harris for providing a definition of the field-star $M_V$,
$V$$-$$I$ main sequence and for unpublished information. We would also like to
thank G.~Jacoby for instructive conversations at the start of this project and
on the properties of optically thin and thick nebulae, and for his
participation as a Co-Investigator on the Cycle~3 portion of the project.
S.~Torres-Peimbert provided support at a critical moment. This work was
supported by STScI grants GO-04308.01-92A and GO-06119.02-94A, and NSF NYI
grant AST 92-577833.
\newpage
|
\section{Introduction}
The rapid expansion of the use of high quality crystaline materials in optical
and eletronic devices has strongly stimulated research, both theoretical and
experimental, on dynamics of crystalization. Computer simulation has played an
important role on the development and understanding of crystal growth. During growth
the solid-fluid interface can display several interesting phenomena like segregation,
dynamical instabilities and pattern formation \cite{cross}. \\
A crystal can grow from the adjacent fluid by different mechanisms, depending on the
structure of the interface (rough or smooth), material purity, growth rates,
temperature gradients and so on. For a crystal to grow atoms or molecules must be
transported from the fluid towards the interface with a non-zero sticking probability.
besides that, the latent heat generated at the interface as well as the solute excess
segregated must be carried away. These requirements can be met in a controlled way by
puting the sample in an appropriated furnace submmited to a temperature gradient and
pulling it with constant speed toward the colder region. results of such experiments
have been compared with results of two dimensional models. Nijmeijr and
Landau\cite{marco} reported molecular dynamic (MD) simulation on laser heated pedestal
growth of fibers. Previous simulations consisted of kinetic models using Monte Carlo
techniques and numerical solution of transport equations\cite{weeks}. \\
\section{Simulation}
In this work we use MD to simulate the solidification of a two component system of a
solvent (particles of type a) and solute (particles of type b) interating via a
Lennard-Jones (LJ) potential. By tunning the LJ parameters we set the struture of the
interface (rough or smooth) and the segregation coefficient. We also simulate
morphological instabilities, the planar inteface becoming cellular and eventually
dendritic (Mullins-sekerka instability.). The simulation is carried out with all
particles interacting through the LJ potential
\begin{equation}
\phi_{i,j}(r_{i,j}) = \epsilon_{i,j}
\left [
(\frac{\sigma_{i,j}}{r_{i,j}})^{12} -
(\frac{\sigma_{i,j}}{r_{i,j}})^{6} .
\right ]
\end{equation}
The indexes $i$ and $j$ stands for particles in the positions $r_i$ and $r_j$
respectively and $r_{i,j} = |r_i - r_j|$. There are three types of interactions,
solvent-solvent, solvent-solute and solute-solute. In each case the LJ parameters are
labeled as $(a,a)$, $(a,b)$ and $(b,b)$. distance $r$, time $t$ and temperature $T$
are measured in units of $\sigma_{a,a}$,
$\sigma_{a,a}(m_a/\epsilon_{a,a})^{\frac{1}{2}}$ and $k_{B}/\epsilon_{a,a}$. initially
we distribute $N=n_x \times n_z$ particles over the two dimensional volume
$L_x \times L_z$. We assume periodic boundary conditions in the $x$ direction. In the
$z$ direction we divide the system in to two distinct regions, a solid and a fluid
one. In the solid region particles stand initially in their equilibrium position in a
total of $n_x \times n_{0z}$ particles. on the fluid region we distribute the rest of
the particles with the density chosen to be $\rho = 0.5\sigma_{a,a}^{-2}$, randomly
distributed in a triangular lattice and slightly dislocated from their equilibrium
position. We impose a temperature gradient along the $z$ direction using a velocity
renormalization approach. The system is divided in two regions, one cold which is set
to $T=0$ and the other to $T=T_h$ higher than the melting temperature $T_m$ which is
$T_m=0.4$ in our units. we let the system evolve in time for $N_e$ time steps of size
$\delta t$ until it reachs equilibrium. Once the equilibrium is reached we start
pulling the system in the $+z$ direction, at a pulling velocity $V_p$. The $-z_{max}$
layer works as a particle source, maintaining a constant flow to the material. The
basic experimental set up is shown in Fig.~\ref{eps1}
\begin{figure}
\includegraphics[width=.9\textwidth]{figure1.eps}
\caption[]{Basic experimental set up for directional growth. The interface motion is
visualized using an optical microscope. $T_m$ is the melting temperature of the
mixture and $V_p$ is the pulling velocity.}
\label{eps1}
\end{figure}
In Fig.~\ref{eps2} we show an example of solute segregation during directional growth
o fthe binary mixture: caprolactane as solvent and methyl-blue as solute. In the top
part of Fig.~\ref{eps2} is shown an image of the crystal (left side) and melt (right
side), with maximum concentration of methyl-blue at the melt side of the interface.
From the gray level of the image we obtain the methyl-blue concentration profile
across the sample, which in the melt, decays exponetially as a function of the
distance from the interface (bottom part of Fig.~\ref{eps2}). Fig.~\ref{eps3} shows
the simulation results done in a system initially with $27 \times 30$ particles,
solute concentration of $5\%$, pulling velocity $V_p = 0.004$, and LJ parameters given
by : $\epsilon_{a,b}=0.5$, $\epsilon_{b,b}=0.1$, $\sigma_{a,b}=1.0$, $\sigma_{b,b}=1.0$
and $m_b = 1.0$.
\begin{figure}
\includegraphics[width=.9\textwidth]{figure2.2.ps}
\caption[]{Directional solidification of a binary mixture showing the solute
segregation at the interface. The experimental result is for the mixture caprolactane
(solvent) and methyl-blue (solute).}
\label{eps2}
\end{figure}
After averaging over many runs to improve statistics one obtains the steady-state
solute concentration profile represented as data points with error bars.
\begin{figure}
\includegraphics[width=.9\textwidth]{figure3.eps}
\caption[]{Our simulation results using the Lennard-Jones potential describes in
detail the experiment.}
\label{eps3}
\end{figure}
It is shown in Fig.~\ref{eps3} a theoretical result obtained from the diffusion
equation for the solute concentration in the fluid phase\cite{coura}
\begin{equation}
\frac{\partial c_f}{\partial t} =
D\frac{\partial^2 c_f}{\partial z^2} + V_f \frac{\partial c_f}{\partial z}
\end{equation}
where $D$ is the solute diffusion coefficient in the fluid phase and $V_f$ is the
velocity of the system of reference. The diffusion equation is supplemented with the
boundary conditions:
\begin{subeqnarray}
c_f & = & c_0 ~~~~~~~~~~~~~~~~ at ~~ z \rightarrow \infty \ts , \label{f1a}\\
(1-K)V_fc_f & = & -D\frac{d c_f}{d z} ~~~~~~~~ at ~~ z = 0 \ts .\label{f1b}
\end{subeqnarray}
The second condition is just the mass conservation at the interface and $K$ is the
ratio between the solute and solvent concentrations.
\section{Cellular Instability}
With the LJ parameters chosen above the morphological instabilities were inhibited
during the cellular growth. We can stimulate instabilities by tunning the LJ
parameters, particularly by slightly varying the equilibrium position between
particles $a$ and $b$, i.e., changing $\sigma_{a,b}$. Fig.~\ref{eps4} show the effect
of varying this parameter.
\begin{figure}
\includegraphics[width=.9\textwidth]{figure4.eps}
\caption[]{The figure shows the concentration profile found in our simulation. It
is to be compared to Fig.~\ref{eps2} }
\label{eps4}
\end{figure}
For $\sigma_{a,b} = 1$ the interface is rough, becoming cellular and dendritic for
$\sigma_{a,b} > 1$. As shouls be expected a lower $\sigma_{a,b}$ sets the interface
smoother than larger values indicating that the system is very sensitive to geometric
factores. (We have also simulated some experiments fixing the value of $\sigma_{a,b}$
and varying $\epsilon_{a,b}$ and $\epsilon_{b,b}$, the results did not show any
qualitative change in their behavior. Raising $\sigma_{a,b}$ breaks the hexagonal
structure o fthe solid. The energy involved in such distortion is so large that the
interface segragate solute particles, bonds $(a,a)$ and $(b,b)$ are preferible than
$(a,b)$. Once particles of the type $b$ concentrate at the interface its melting point
is lowered and channels of $b$ particles are formed inside the crystal. Fig.~\ref{eps5}
Shows this effect.
\begin{figure}
\includegraphics[width=.8\textwidth]{figure5.2.ps}
\caption[]{Cellular instability originated by varying
$\sigma_{a,b}$. From left to right
$\sigma_{a,b}=1.05,1.10$ and $1.20.$}
\label{eps5}
\end{figure}
We are now studying the stability \cite{figueiredo} of those interfaces. We start with
a smoth interface then, we introduce a perturbation at the interface. It has been
shown that near the bifurcation from planar to cellular the time evolution of the most
instable Fourier mode of the perturbation can be described by a third order
Landau-amplitude equation. This however is the subject of a new research and will be
soon published elsewhere.
\section{Acknowledgements}
This work was partially supported by CNPq, FAPEMIG and FINEP. Numerical work was done
at the CENAPAD-MG/CO and in the LINUX parallel cluster at the {\sl Laborat\'orio de
Simula\c{c}\~ao} Departamento de F\'{\i}sica - UFMG.
|
\section{Introduction}
The main goal of this paper is to demonstrate, on the examples of
semisimple algebras of second order ($ A_2, B_2, C_2, G_2$),
the general
construction connecting a semisimple algebra of a given grading to an
exactly integrable system.
The simplest example is the two-di\-men\-sional Toda lattice considered and
integrated in the case of an arbitrary semisimple algebra almost 20 years ago
\cite{leznov:LG, leznov:ls0}\footnote{For $A_n$ series this problem was solved more then 150 years
ago in Darbouxs papers!}. For
main grading, exactly integrable systems were explicitly found and described in
the recent papers of the author \cite{leznov:LM} (so called Abelian~case).
In the present paper we follow three dif\/ferent and independent aims.
The f\/irst is to relate unknown up to now integrable systems to
nonabelian gradings\footnote{Note that the zero order subspace is a non-commutative
algebra by itself.} (see \cite{leznov:GC} in this con\-nec\-tion).
The second one is to get rid of the restriction of nonabelian Toda theory
to use only subspaces with zero and $\pm 1$ grading indices. The last, but not
the least important one, is to provide the reader with a scheme of how the group
representation theory (in the very restricted volume) can be applied to the
theory of integrable systems.
For our purposes here it is not the shortest and simplest way to the result that
is important, but the result by itself. Therefore, in concrete examples we
tried to use calculations that can be followed and checked directly using only
simplest algebra.
The paper is organized as follows. Section~2 contains the background information
on representation theory of semisimple algebras and groups (as a rule without
proofs). Section~3 describes the general construction, mathematical tricks and
methods used in main sections. In Section~4 concrete examples of semisimple
algebras of second order are considered in details for all possible
gradings. Concluding remarks and perspectives for further investigation
are outlined in Section~5.
\section{Semisimple algebras and groups}
Let ${\cal G}$ be an arbitrary f\/inite-dimensional graded Lie
algebra\footnote{We make no distinction between algebras and super-algebras,
just keeping in mind that even (odd) elements of super-algebras are always
multiplied by even (odd) elements of the Grassman space.}.
Then $\cal G$ can be written as a direct sum of subspaces of dif\/ferent grading
indices
\begin{equation}
{\cal G}=\left(\oplus^{N_-}_{k=1} {\cal G}_{-\frac{k}{2}}\right)
{\cal G}_0 \left(\oplus^{N_+}_{k=1}{\cal G}_{\frac{k}{2}}\right).
\label{GR}
\end{equation}
Generators with an integer grading index are called bosonic, while those with
half-integer grading index are named fermionic. The positive (negative) grading
corresponds to upper (lower) triangular matrices.
The grading operator $H$ for an arbitrary semisimple algebra can be written
as a linear combination of elements of commutative Cartan subalgebra taking
unity or zero values on the generators of simple roots
\begin{equation}
H=\sum^r_{i=1} \left(K^{-1}c\right)_i h_{i}.
\label{cartan1}
\end{equation}
Here $K^{-1}$ is the inverse Cartan matrix $K^{-1}K=KK^{-1}=I$
and $c$ is a column of zeros and unites in an arbitrary order.
Under the main grading all $c_i=1$. In this case
$(K^{-1}c)_i=\sum\limits_{j=1}^r K^{-1}_{i,j}$, where $r$ is the rank of the algebra.
As usually, generators of simple roots $X^{\pm}_i$ (raising/lowering
operators) and Cartan elements $h_i$ satisfy the system of commutation
relations:
\begin{equation}
[h_i , h_j]=0, \qquad [h_i,X^{\pm}_j]=\pm K_{j,i}X^{\pm}_j, \qquad
[X^{+}_i,X^{-}_j\}={\delta}_{i,j} h_j, \quad (1 \leq i,j \leq r),\label{aa6}
\end{equation}
where $K_{ij}$ are elements of Cartan matrix and brackets $[,\}$
stand for the graded commutator.
The highest vector $\ve{j}$ ($\vc{j} \equiv \ve{j}^{\dagger}$) of
the $j$--th fundamental representation has the following properties:
\begin{equation}
X^{+}_i\ve{j}=0, \qquad h_i\ve{j}={\delta}_{i,j}\ve{j}, \qquad \vc{j}\ve{j}=1.
\label{high}
\end{equation}
The representation is exhibited by applying
lowering operators $X^{-}_i$ to the vector $\ve{j}$ repeatedly and
extracting all linearly-independent vectors with non-zero norm. The f\/irst
few basis vectors are
\begin{equation}
\ve{j}, \quad X^{-}_j\ve{j}, \quad X^{-}_i X^{-}_j\ve{j}\neq 0,\quad
K_{i,j}\neq 0,\quad i\neq j.
\label{vectors}
\end{equation}
}
In fundamental representations an important identity for matrix elements of a
group element $G$ holds\footnote{Recall that a
superdeterminant is def\/ined as $\mbox{sdet} \left(\begin{array}{cc} A, & B \\ C, & D
\end{array}\right) \equiv \det (A-BD^{-1}C ) (\det D)^{-1}$.}~\cite{leznov:ls0}
\begin{equation}
\mbox{sdet} \left(\begin{array}{cc} \vc{j}X_j^+GX_j^-\ve{j}, &
\vc{j}X_j^+G\ve{j}
\vspace{3mm}\\
\vc{j}GX_j^-\ve{j}, & \vc{j} G \ve{j} \end{array}\right) = \prod^r_{i=1}\vc{i} G \ve{i}^{-K_{ji}},
\label{recrel}
\end{equation}
The identity (\ref{recrel}) is in fact a generalization (to the case of an arbitrary semisimple
Lee super-group) of the famous Jacobi identity that relates
determinants of orders $(n-1)$, $n$ and $(n+1)$ of some special
matrices. As we will see in the next section, this identity is of
such importance in deriving exactly integrable systems that one
can even say that it is responsible for their existence. We will
still refer to (\ref{recrel}) as to ``the f\/irst Jacobi
identity''. In addition to (\ref{recrel}), there exists another
independent identity of key importance~\cite{leznov:l}
\begin{equation}
\hspace*{-5pt}\begin{array}{l} \displaystyle (-1)^P K_{i,j} {\vc{j}X_j^+X_i^+ G \ve{j}\over \vc{j} G
\ve{j}}+ K_{j,i} {\vc{i}X_i^+X_j^+ G \ve{i}\over \vc{i} G \ve{i}}
\vspace{3mm}\\ \displaystyle \qquad +K_{ij}K_{j,i}(-1)^{jP}{\vc{j}X_j^+ G
\ve{j}\over \vc{j} G \ve{j}}{\vc{i}X_i^+ G \ve{i}\over \vc{i} G
\ve{i}}=0 ,\qquad i\neq j \end{array}\label{J2} \end{equation} which will be called
the second Jacobi identity. This identity is responsible (in the
above sense) for the existence of hierarchies of integrable
systems invariant with respect to integrable mappings that are
connected to every exactly integrable system.
Either from (\ref{recrel}) or from (\ref{J2}) it is possible to construct
many usefull recurrent relations that are used in further
consideration.
Taking into account the importance of Jacobi identities (\ref{recrel})
and (\ref{J2}) for further consideration we
present below a brief proof of (\ref{recrel}).
Let us consider the left hand side of (\ref{recrel}) as a function on the
group. The action on an arbitrary group element $G$ in the def\/inite
representation $l$ of the operators of the right (left) regular
representation is by def\/inition
\begin{equation}
M_{\mbox{\scriptsize left}}(\tilde M_{\mbox{\scriptsize right}}) G= M_l G (\tilde M_l),
\label{AM}
\end{equation}
where $M_l$, $\tilde M_l$ are the generators (the matrices of
corresponding dimension) of shifts on the group in a given $l$
representation. Now let us act with an arbitrary generator of the
simple positive root $(X^+_s)_r$ on the left hand side of
(\ref{recrel}). This action is equivalent to dif\/ferentiation and
therefore should be applied consequently to the f\/irst and second
columns of the matrix (\ref{recrel}) adding the results. The
action on the second column results in zero as a corollary of the
def\/inition of the higest state vector (\ref{high}). Action on
the f\/irst column is dif\/ferent from zero only in the case
$s=j$. But in this case using the same def\/inition of the highest
state vector we conclude that as a result of dif\/ferentiation of
the f\/irst column it becomes equal to the second one with the
zero f\/inal result. Thus considered as a function on the group
the left hand side of (\ref{recrel}) is also proportional to the
highest vector (or a linear combination of such vectors) of some
other representation. The higest vector of the irreducible
representation is uniquely def\/ined by the values that Cartan
generators take on it. If Cartan generators take on the highest
vector values $V(h_i)=l_i$, the last can be uniquely represented
in the form
\begin{equation}
\vc{l} G \ve{l}=C \prod_{i=1}^r(\vc{i} G \ve{i})^{l_i}.\label{HVD}
\end{equation}
Calculating the values of Cartan generators on
the left hand side of equation (\ref{recrel}) (both left and right with
the same result) and using the last comment
about the form of the highest vector, we prove (\ref{recrel}) ($C=1$, as can be
seen by putting $G=1$ and comparing both sides).
The second Jacobi identity can be proven by similar argument \cite{leznov:l}.
The following generalization of the f\/irst Jacobi identity will be very
important in calculations dealing with nonabelian gradings.
Let $\ve{\alpha}$ be basis vectors of some representation in the
strict order of increasing the number of lowering generators (see
(\ref{high}) and (\ref{vectors})). We also assume that the action
of a generator of an arbitrary positive simple root on each basis
vector results in a linear combination of the previous ones.
Then the principal minors of an arbitrary order of the matrix
($G$ is an arbitrary element of the group):
\[
G_{\alpha}=\vc{\alpha} G \ve{\alpha}
\]
are annihilated from the right (from the left) by generators of positive
(negative) roots.
Indeed this is equivalent to dif\/ferentiation and therefore it is
necessary to act on each column (line) of the minors matrix and
add the results. But the action of the generator of a positive
simple root on the state vector with a given number of lowering
operators transforms it into a state vector with a number of
lowering operators on unity less, which according to our
assumption is a linear combination of previous columns (lines).
Thus in all cases the lines or columns of the resulting
determinant are linearly dependent with zero result.
The generators of Cartan subalgebra obviously take the def\/inite values on
minors of these kind and if the corresponding values are
$l^s_i$, it is possible to write the equality in correspondence with (\ref{HVD})
\begin{equation}
Min_s= C_s \prod_{i=1}^r \vc{i} G \ve{i}^{l^s_i},\label{GJI}
\end{equation}
where constants $C_s$ can be determined as described above.
\setcounter{equation}{0}
\section{General construction and technique of computation}
The grading of a semisimple algebra is def\/ined by the values that the grading
operator $H$ takes on the simple roots of the algebra. As it was mentioned
above, this values can be only zeros and unites in an arbitrary order.
\[
[H, X^{\pm}_i]=\pm X^{\pm}_i,\qquad H=\sum_1^r (K^{-1}c)_i h_i,\qquad c_i=1,0.
\]
On the level of Dynkin's diagrams the grading can be introduced by
using two colors for its dots: black for simple roots with $c_i=1$
and red for roots with $c_i=0$. To each consequent sequence of the
red (simple) roots the corresponding semisimple algebra
(subalgebra of the initial one) is connected. All these algebras
are obviously mutually commutative and belong to the zero graded
subspace. Cartan elements of the black roots also belong to the
zero graded subspace. We will use the usual numeration of the dots
of Dynkin diagrams and all red algebras will be distinguished by
an index of their f\/irst root~$m_s$. The rank of $m_s$-th red
algebra will be denoted as $R_s$. Thus
$X^{\pm}_{m_s},X^{\pm}_{m_s+1},\ldots,X^{\pm}_{m_s+R_s-1}$ is the
system of simple roots of $m_s$ red algebra.
After these preliminary comments turn to the general
construction \cite{leznov:l1}.
Let two group valued functions $ M^+(y)$, $M^-(x)$ be solutions of $S$-matrix
type equations
\begin{equation}
\hspace*{-5pt}\begin{array}{l}
\displaystyle M^+_y=\left(\sum_0^{m_2} B^{(+s}(y)\right) M^+\equiv\left(B^{(0}+L^+\right)M^+,
\vspace{3mm}\\
\displaystyle M^-_x=M^-\left(\sum_0^{m_1} A^{(-s}(x)\right)\equiv M^-\left(A^{(0}+L^-\right),
\end{array}
\label{I}
\end{equation}
where $B^{(+s}(y)$, $A^{(-s}(x)$ take values in $\pm s$ graded
subspaces correspondingly and $s=0, 1, 2,\ldots,m_{1,2}$. In each f\/inite-dimensional
representation $B^{(+s}(y)$, $A^{(-s}(x)$ are upper (lower) triangular matrices and therefore
equations~(\ref{I}) are integrated in quadratures.
The composite group valued function $K$ plays the key role in our construction
\begin{equation}
K=M^+ M^-. \label{II}
\end{equation}
It turns out that matrix elements of $K$ in various fundamental
representations are related by closed systems of equivalent relations,
which can be interpreted as exactly integrable system with known general
solution.
Bellow we describe calculation methods to prove
this proposition.
First of all let us calculate the second mixed derivative $(\ln \vc{i} K
\ve{i})_{x,y}$, where index $i$ belongs to the black dot of Dynkin diagram.
We have
\begin{equation}
(\ln \vc{i} K \ve{i})_x= {\vc{i} K (A^0+L^-) \ve{i}\over \vc{i} K \ve{i}}=
A^0_i(x)+ {\vc{i} K L^- \ve{i}\over \vc{i} K \ve{i}}.
\label{MI}
\end{equation}
Indeed, $ K_x=M^+(y)M^-_x(x)=K (A^0+L^-)$ as a corollary of equation for
$M^-$. All red components of $A^0$ under the action on the black highest
vector state $ \ve{i}$ lead to zero result in connection with
(\ref{vectors}).
The action of Cartan elements of the black roots state vector satisf\/ies
the condition $h_j \ve{i}=\delta_{i,j} \ve{i}$ and thus only coef\/f\/icient on
$h_i$ remains in the f\/inal result (\ref{MI}).
Further dif\/ferentiation (\ref{MI}) with respect to $y$, with the
help of arguments above, leads to following result:
\begin{equation}
(\ln \vc{i} K \ve{i})_{x,y}=\vc{i} K \ve{i})^{-2} \pmatrix{
\vc{i} K \ve{i}, & \vc{i} K L^- \ve{i} \cr
\vc{i} L^+ K \ve{i}, & \vc{i} L^+ K L^- \ve{i} \cr}.
\label{AR}
\end{equation}
Applying (\ref{AM}) of the previous section to the left hand side of (\ref{AR}), we f\/inally obtain
\begin{equation}
(\ln \vc{i} K \ve{i})_{x,y}=L^-_rL^+_l \ln \vc{i} K \ve{i})^{-1}.
\label{ARR}
\end{equation}
Thus the problem of calculating the mixed second derivative is
reduced to purely algebraic manipulations on the level of representation theory
of semisimple algebras and groups. Further evaluation of (\ref{ARR}) is
connected with repeated application of the f\/irst~(\ref{recrel})
and second (\ref{J2}) Jacobi identities as it will be clear from the material
of the next section.
As it was mentioned above, the red algebras of zero order graded subspace in
general case are not commutative. This leads to additional
computational dif\/f\/iculties. Let us denote by $ \ve{m_i}$
the highest vector of $m_i$th fundamental representation of the initial
algebra. Of course, $\ve{m_i}$ is simultaneously the highest vector of the
f\/irst fundamental representation of the $m_i$ red algebra. Let $\vc{\alpha_i},
\ve{\beta_i}$ be basis vectors of the f\/irst fundamental representation
(this restriction is not essential) of $m_i$-th red algebra and let us
consider the matrix elements of element $K$ in this basis.
$R_i+1\times R_i+1$ matrix ($R_i+1$ is the dimension of the f\/irst fundamental
representation) with matrix elements $\vc{\alpha_i} K \ve{\beta_i}$
will be denoted by a single symbol $u_i$ (index $i$ takes values from one to
the number of the red algebras, which is the function of the choosen grading).
For derivatives of matrix elements of so constructed matrix we have
consequently (index~$i$ we omite for a moment):
\begin{equation}
\vc{\alpha} u_x \ve{\beta}=\vc{\alpha} K (A^0+L^-) \ve{\beta}=\! \sum_
{\gamma} \vc{\alpha} K \ve{\gamma}\vc{\gamma} I A^0 \ve{\beta}
+\vc{\alpha} K L^- \ve{\beta}. \label{MCD}
\end{equation}
Or equivalently
\[
u^{-1} u_x= A^0(x)+u^{-1} \vc{} K L^- \ve{}.
\]
Further dif\/ferentiation with respect to $y$ variable leads to
\begin{equation}
\hspace*{-5pt}\begin{array}{l}
\displaystyle \vc{}((u^{-1} u_x)_y\ve{}=u^{-1} \vc{} (B^0+L^+) K L^- \ve{}-u^{-1}\vc{}
(B^0+L^+) K \ve{} u^{-1}\vc{} K L^-\ve{}
\vspace{2mm}\\
\displaystyle \qquad = u^{-1}(\vc{} L^+ K L^- \ve{}-\vc{} L^+ K \ve{} u^{-1} \vc{} K L^-\ve{}).
\end{array} \label{MC}
\end{equation}
The last expression may be brought to the form of the ratio of
two determinants of $R_i+2$ and $R_i+1$ orders respectively with
the help of standard transformations:
\begin{equation}
\vc{} u (u^{-1} u_x)_y \ve{}={\mbox{Det}_{N_i+1}\pmatrix{ u & K L^-\ve{} \cr
\vc{} L^+ K & \vc{} L^+ K L^- \ve{} \cr}\over \mbox{Det}_{N_i}(u)}. \label{MCC}
\end{equation}
The generalised Jacobi identity (\ref{GJI}) of the previous section plays the key role for
discovery of the last expression and will be exploited many times.
\setcounter{equation}{0}
\section{The algebras of second rank {\mathversion{bold}$A_2$, $B_2, C_2$, $G_2$}}
All elements of these algebras may be constructed by consequent
multi-commutation of generators of four simple roots $X^{\pm}_{1,2}$ with the
basic system of commutation relations
\begin{equation}\label{RS}
\hspace*{-5pt}\begin{array}{l}
[X^+_1,X^-_1]=h_1,\qquad [X^+_1,X^-_2]=[X^+_2,X^-_1]=0,\qquad[X^+_2,X^-_2]=h_2,
\vspace{2mm}\\
\displaystyle {}[h_1,X^{\pm}_1]= {\pm} 2X^{\pm}_1, \qquad [h_2,X^{\pm}_2]= {\pm} 2X^{\pm}_2,
\vspace{2mm}\\
\displaystyle {}[h_1,X^{\pm}_2]= {\mp} pX^{\pm}_2, \qquad [h_2,X^{\pm}_1]= {\mp} X^{\pm}_1,
\qquad p=1,2,3.
\end{array}
\end{equation}
In all cases there are three possible nontrivial gradings: $(1,1)$ -- the
principle one (Abelian case), $(1,0)$ -- the grading of the f\/irst simple root
and $(0,1)$ -- of the second simple one. In the case of the principle grading
corresponding integrable systems for arbitrary semisimple algebras were found
and described in \cite{leznov:LM}. Each further subsections will be
devoted to detail consideration of nonabelian gradings $(1,0)$, $(0,1)$, which
are equivalent to each other only in the case of $A_2$ algebra.
In the end of this mini-introduction we present the second Jacobi identity
as applied to the algebras of second rank:
\begin{equation}
{\vc{2} X^+_2 X^+_1 K \ve{2}\over \vc{2} K \ve{2}}+p {\vc{1} X^+_1 X^+_2 K
\ve{1}\over \vc{1} K \ve{1}}=p {\vc{2} X^+_2 K \ve{2}\over \vc{2} K \ve{2}}
{\vc{1} X^+_1 K \ve{1}\over \vc{1} K \ve{1}}\label{2JI}
\end{equation}
or in notation, which will be introduced by the way of consideration
\[
\bar \alpha_{21}+p\bar \alpha _{12}=p\bar \alpha _1\bar \alpha _2,\qquad
\alpha_{12}+p\alpha _{21}=p\alpha _1 \alpha _2.
\]
\subsection{Unitary {\mathversion{bold}$A_2$} serie}
The root system of this algebra consists of three elements with the
generators $X^{\pm}_1,X^{\pm}_1,X^{\pm}_{12}$ $\equiv \pm [X^{\pm}_1,X^{\pm}_2]$.
This case corresponds to $p=1$ in (\ref{RS}). For def\/initeness we restrict
ourselves by $(1,0)$ grading $[H,X^{\pm}_1]=\mp X^{\pm}_1$, $[H,X^{\pm}_2]=0$.
$L^{\pm}$ operators belong to $\pm 1$ graded subspaces and have the form:
\[
L^+=\bar c_1 X^+_1+\bar c_2 [X^+_2,X^+_1],\qquad L^-=c_1 X^-_1+c_2 [X^-_1,X^-_2],
\]
where $c_{1,2}\equiv c_{1,2}(x)$, $\bar c_{1,2}\equiv \bar c_{1,2}(y)$.
The object of investigation is $2\times 2$ matrix $u$ in the basis of the
second fundamental representation of $A_2$ algebra\footnote{The (bra) basis
vectors of the three dimensional (``qwark'') second fundamental representation
of $A_2$ algebra are the $\vc{2}$, $\vc{2} X^+_2$, $\vc{2} X^+_2 X^+_1 $.}:
\begin{equation}
u=\left(\begin{array}{cc} \vc{2} K \ve{2}, & \vc{2} K X_2^-\ve{2} \vspace{1mm}\\
\vc{2}X_2^+ K \ve{2}, & \vc{2}X_2^+ K X_2^-\ve{2}
\end{array}\right). \label{recrelI}
\end{equation}
In correspondence with (\ref{MCC}) we have:
\begin{equation}
\vc{} u (u^{-1} u_x)_y \ve{}={\mbox{Det}_3\pmatrix{ u & I K L^- \ve{} \cr
\vc{} L^+ K I & \vc{} L^+ K L^- \ve{} \cr}\over \mbox{Det}_2(u)}.
\label{MCC"}
\end{equation}
The action of operators $L^{\pm}$ on basis vectors $\ve{2}, X^-_2 \ve{2}$
($\vc{2},\vc{2} X^+_2$) is the following:
\[
\hspace*{-5pt}\begin{array}{l}
L^-\ve{2}=c_2 X^-_1 X^-_2 \ve{2}, \qquad L^-X^-_2\ve{2}=c_1 X^-_1 X^-_2 \ve{2},
\vspace{2mm}\\
\vc{2}L^+=\bar c_2 \vc{2} X^+_2 X^+_1,\qquad \vc{2} X^+_2 L^+=\bar c_1
\vc{2} X^+_2 X^+_1.
\end{array}
\]
So in this case the following sequence of basis vectors from generalized
Jacobi identity (\ref{GJI}) takes places:
\[
\vc{2}, \quad \vc{2} X^+_2, \quad \vc{2} X^+_2 X^+_1.
\]
The summed values of Cartan generators $h_1$, $h_2$ on this basis take
zero values and so $\mbox{Det}_3$ from (\ref{MCC"}) equal to unity (with correct
account of the constant). This is a really highest vector of scalar,
one-dimensional representation of $A_2$ algebra.
Finally (\ref{MCC"}) leads to the system, which matrix function $u$ satisfy:
\begin{equation}
(u^{-1} u_x)_y =(\mbox{Det}\; u)^{-1} u^{-1}\pmatrix{ c_2 \bar c_2, & c_1 \bar c_2 \cr
c_2 \bar c_1, & c_2 \bar c_2 \cr}.
\label{A_2}
\end{equation}
In usual notations the system (\ref{A_2}) is nonabelian $A_2 (1,0)$ Toda
chain. The system (\ref{A_2}) is obviously form-invariant with respect to
transformation:
\[
u\to \bar g(y) u \bar g(x).
\]
With the help of this transformation the arbitrary up to now functions
$c$, $\bar c$ may be evaluated to a constant values.
\subsection{Orthogonal {\mathversion{bold}$B_2$} serie equivalent to simplectic one
{\mathversion{bold}$C_2$}}
This case corresponds to the choise $p=2$ in (\ref{RS}). Both
gradings are not equivalent to each other and must be considered
separately. First fundamental representation for $B_2$ algebra is
the second one for $C_2$ serie and vice versa.
\subsubsection{(1,0) grading}
Generators $L^{\pm}$ may contain components with $\pm 1$, $\pm2$ graded
indexes and have the form:
\[
\hspace*{-5pt}\begin{array}{l}
L^+=\bar c_1 X^+_1+\bar c_2 [X^+_2,X^+_1]+\bar c^2 [[X^+_2,X^+_1]X^+_1],
\vspace{2mm}\\
L^-=c_1 X^-_1+c_2[X^-_1,X^-_2]+c^2 [ X^-_1[X^-_1,X^-_2]].
\end{array}
\]
The object of investigation is two dimensional matrix $u$ in the basis of
the second fundamental representation of $B_2$ algebra. The main equation
(\ref{MCC"}) also does not change. The action of $L^{\pm}$ operators on
the basis vectors have now the form\footnote{Five basis vectors of the f\/irst fundamental
representation of the $B_2$ algebra are the following: $ \ve{2}$, $X^-_2 \ve{2}$,
$X^-_1 X^-_2 \ve{2}$, $X^-_1 X^-_1 X^-_2 \ve{2}$, $X^-_2 X^-_1 X^-_1 X^-_2 \ve{2}$.}:
\[
\hspace*{-5pt}\begin{array}{l}
L^-\ve{2}=(c_2+c^2 X^-_1) X^-_1 X^-_2 \ve{2}, \qquad L^-X^-_2\ve{2}=
(c_1 +c^2 X^-_2 X^-_1) X^-_1 X^-_2 \ve{2},
\vspace{2mm}\\
\vc{2}L^+=\vc{2} X^+_2 X^+_1(\bar c_2+\bar c^2 X^+_1),\qquad
\vc{2} X^+_2 L^+= \vc{2} X^+_2 X^+_1 (\bar c_1+\bar c^2 X^+_1 X^+_2).
\end{array}
\]
Substituting this expression into (\ref{MCC"}) after some trivial
evaluations we come to the following relation:
\begin{equation}
\hspace*{-5pt}\begin{array}{l}
\displaystyle u(u^{-1} u_x)_y =(\mbox{Det}\; u)^{-1}
\pmatrix{ \bar c_2+\bar c^2 (X^+_1)_l, & 0 \cr
\bar c_1+\bar c^2 (X^+_1 X^+_2)_l , & 0 \cr}
\vspace{3mm}\\
\displaystyle \qquad \qquad \qquad \times \pmatrix{ c_2+c^2 (X^-_1)_r , & c_1 +c^2 (X^-_2 X^-_1)_r \cr
0 , & 0 \cr} \mbox{Det}_3.
\end{array} \label{F}
\end{equation}
In the last expression $\mbox{Det}_3$ satisfy all conditions of (\ref{GJI}),
with the sequence of bases vectors:
\[
\vc{2}, \quad \vc{2} X^+_2, \quad \vc{2} X^+_2 X^+_1.
\]
In this case the summed value of Cartan element $h_1$ is equal to
2, of $h_2$ -- to 0. So with the correct value of numerical factor we
obtain $\mbox{Det}_3=2 \vc{1} K \ve{1}^2 $.
The action of the f\/irst line operator in (\ref{F}) on $(\vc{1} K \ve{1})^2$
leads to the line of the form:
\begin{equation}
2 (\vc{1} K \ve{1})^2 (c_2+2 c^2 \alpha_1 , c_1 +2 c^2 \alpha_{21}) \label{L},
\end{equation}
where following abbreviations are used:
\begin{equation}
\hspace*{-5pt}\begin{array}{l}
\displaystyle \bar \alpha_1={\vc{i} X^+_i K \ve{i}\over \vc{i} K \ve{i}},\qquad
\bar \alpha_{12}={\vc{1} X^+_1 X^+_2 K \ve{1}\over \vc{1} K \ve{1}},
\vspace{3mm}\\
\displaystyle \bar \alpha_{21}={\vc{2} X^+_2 X^+_1 K \ve{2}\over \vc{2} K \ve{2}}, \qquad
\alpha_i={\vc{i} K X^-_i \ve{i}\over \vc{i} K \ve{i}},\qquad i=1,2,
\vspace{3mm}\\
\displaystyle \alpha_{21}={\vc{1} K X^-_2 X^-_1 \ve{1}\over \vc{1} K \ve{1}},\qquad
\alpha_{12}={\vc{2} K X^-_1 X^-_2 \ve{2}\over \vc{2} K \ve{2}}.
\end{array}\label{NOT}
\end{equation}
Now it is necessary to act with the help of the column operator (\ref{F})
on the line (\ref{L}). The result of this action on scalar factor may be
presented in the form ($\mbox{Det}_2 u=\vc{1} K \ve{1}^2$):
\[
2\pmatrix{ \bar c_2+2 \bar c^2 \bar \alpha_1, & 0 \cr
\bar c_1+2 \bar c^2 \bar \alpha_{12} , & 0 \cr}
\pmatrix{ c_2+2 c^2 \alpha_1 , & c_1 +2 c^2 \alpha_{21} \cr
0 , & 0 \cr}.
\]
The action of the column operator (\ref{F}) on the line (\ref{L}) leads
to additional matrix:
\[
4c^2 \bar c^2 \pmatrix{ (X^+_1)_l \alpha_1 & (X^+_1)_l \alpha_{21} \cr
(X^+_2 X^+_1)_l \alpha_1 & (X^+_2 X^+_1)_l \alpha_{21} \cr}.
\]
With the help of formulae of Appendix~I the last matrix may be evaluated to
the form:
\[
4 c^2 \bar c^2 (\mbox{Det}\; u)^{-1} u .
\]
Gathering all results together, we obtain f\/inally:
\begin{equation}
u(u^{-1} u_x)_y =2 \pmatrix{ p_1 \bar p_1, & p_2 \bar p_1 \cr
p_1 \bar p_2, & p_2 \bar p_2 \cr}+
4 c^2 \bar c^2 (\mbox{Det} \; u)^{-1} u, \label{B_2}
\end{equation}
where
\[
p_1=c_2+2 c^2 \alpha_1,\qquad \bar p_1=\bar c_2+2 \bar c^2 \bar \alpha_1,\qquad
p_2=c_1+2 c^2 \alpha_{21},\qquad \bar p_2=\bar c_1+2 \bar c^2 \bar \alpha_{12}.
\]
Now we would like to show that the derivatives $(p_{\alpha})_y$ and
$(\bar p_{\alpha})_x$ are functionally dependent on matrix $u$ and
themselves, closing in this way the system of equations of equivalence and
representing it in the form of closed system of equations for $8$ unknown
functions: 4~matrix elements of $u$ and $4$ components of $2$ two-dimensional
spinors $p$, $\bar p$.
Let us follow now the main steps of the necessary calculations.
Using the introduced above technique we have subsequently:
\[
(p_1)_y=2 c^2 (\alpha_1)_y=
{ 2 c^2\over \mbox{Det}\, (u)}\;
\mbox{Det}\pmatrix{ \vc{1} K \ve{1} & \vc{1} K X^-_1\ve{1} \cr
\vc{1} L^+ K \ve{1} & \vc{1} L^+ K X^-_1 \ve{1} \cr}.
\]
The action of $L^+$ on the state vector $\vc{1}$ is the following:
\[
\vc{1}L^+=\vc{1} X^+_1(\bar c_1-\bar c_2 X^+_2-2 \bar c^2 X^+_2 X^+_1).
\]
Substituting the last expression in the previous equation and
using the f\/irst Jacobi identity for its two f\/irst terms
(linear in $\bar c_1$, $\bar c_2$) we obtain:
\begin{equation}
\hspace*{-5pt}\begin{array}{l}
\displaystyle (p_1)_y={ 2 c^2\over \mbox{Det}\,(u)}(\bar c_1 \vc{2} K \ve{2} - \bar c_2 \vc{2} X^+_2
K \ve{2})
\vspace{3mm}\\
\displaystyle \qquad -{ 2 c^2\over \mbox{Det}\, (u)} \; \mbox{Det}
\pmatrix{ \vc{1} K \ve{1} & \vc{1} K X^-_1\ve{1} \cr
\vc{1}X^+_1X^+_2 X^+_1 K \ve{1} & \vc{1}X^+_1X^+_2 X^+_1 K X^-_1 \ve{1} \cr}.
\end{array}
\label{AE}
\end{equation}
Substituting into the second Jacobi identity (\ref{2JI}) ($p=2$) the f\/irst
one in the form:
\[
\vc{2} K \ve{2}=\mbox{Det}\pmatrix{ \vc{1} K \ve{1} & \vc{1} K X^-_1\ve{1} \cr
\vc{1} X^+_1 K \ve{1} & \vc{1} X^+_1 K X^-_1 \ve{1} \cr}
\]
we obtain after some trivial transformations equality for two
second order determinants:
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle \pmatrix{ \vc{1} X^+_1 K \ve{1} & \vc{1} X^+_1 K X^-_1\ve{1} \cr
\vc{1}X^+_1X^+_2 K \ve{1} & \vc{1}X^+_1X^+_2 K X^-_1 \ve{1} \cr}
\vspace{3mm}\\
\qquad = \pmatrix{ \vc{1} K \ve{1} & \vc{1} K X^-_1\ve{1} \cr
\vc{1}X^+_1X^+_2 X^+_1 K \ve{1} & \vc{1}X^+_1X^+_2 X^+_1 K X^-_1 \ve{1} \cr}.
\end{array}
\]
Evaluating the last column of the f\/irst determinant with the help of
the f\/irst Jacobi identity:
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle \vc{1} X^+_1 K X^-_1\ve{1}={\vc{2} K \ve{2}+\vc{1} X^+_1 K \ve{1} \vc{1} K
X^-_1\ve{1}\over \vc{1} K \ve{1}},
\vspace{3mm}\\
\displaystyle \vc{1} X^+_1 X^+_2 K X^-_1\ve{1}={\vc{2} X^+_2 K \ve{2}+\vc{1} X^+_1 X^+_2 K
\ve{1} \vc{1} K X^-_1\ve{1}\over \vc{1} K \ve{1}}
\end{array}
\]
we obtain for it:
\[
\bar \alpha_1 \vc{2} X^+_2 K \ve{2}-\bar \alpha_{12} \vc{2} K \ve{2}.
\]
Finally we have:
\begin{equation}
(p_1)_y={ 2 c^2\over \mbox{Det}\,(u)}(u_{11}\bar p_2 -u_{21}\bar p_1),\qquad
(p_2)_y={ 2 c^2\over \mbox{Det}\,(u)}(u_{12 }\bar p_2 -u_{22}\bar p_1). \label{UB_2}).
\end{equation}
So (\ref{B_2}), (\ref{UB_2}) and the same system for derivatives of $(\bar p)_x$
is the closed system of identities or $B_2(1,0;2,2;c^2,\bar c^2)$
exactly integrable system connected with the $B_2$ semisimple serie.
To the best of our knowledge this system was not mentioned in literature
before.
>From the physical point of view the exactly integrable system (\ref{B_2}),
(\ref{UB_2}) may be considered as a model of interacting charge ${1\over 2}$
particle $(\bar p, p)$ with scalar-vector neutral f\/ield $u$.
Putting $ c^2=\bar c^2=0$, we come back to nonabelian Toda lattice system for
single matrix valued unknown function $u$.
\subsubsection{(0,1) grading}
Generators $L^{\pm}$ contain only the components with $\pm 1$ graded
indexes and have the form:
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle L^+=\bar d_1 X^+_2+\bar d_2 [X^+_1,X^+_2]+{1\over2} \bar d_3 [X^+_1[X^+_1,
X^+_2]],
\vspace{3mm}\\
\displaystyle L^-=d_1 X^-_2+d_2 [X^-_2,X^-_1]+{1\over 2} d_3 [ X^-_1[X^-_1,X^-_2]].
\end{array}
\]
With respect to transformation of $1$ -- red group $A_1$ functions
$d_i(x)$, $(\bar d_i(y))$ are components of three dimensional $A_1$ vectors.
The object of investigation is two dimensional matrix $u$ in the basis of
the second fundamental representation of $B_2$ algebra. The main equation
(\ref{MCC"}) conserves its form. The action of $L^{\pm}$ operators on the
basis vectors have now the form\footnote{Four basis vectors of the f\/irst fundamental of
the $C_2$ algebra are the following: $ \ve{1}$, $X^-_1 \ve{1}$,
$X^-_2 X^-_1 \ve{1}$, $X^-_1 X^-_2 X^-_1 \ve{1}$.}:
\[
\hspace*{-5pt}\begin{array}{l}
L^-\ve{1}=(d_2-d_3 X^-_1) , \qquad L^-X^-_1\ve{1}=
(d_1 -d_2 X^-_1) X^-_2 X^-_1 \ve{1},
\vspace{2mm}\\
\vc{2}L^+=\vc{2} X^+_2 X^+_1(\bar d_2-\bar d_3 X^+_1),\qquad
\vc{2} X^+_2 L^+= \vc{2} X^+_2 X^+_1 (\bar d_1-\bar d_2 X^+_1 X^+_2).
\end{array}
\]
Substituting this expression into (\ref{MCC"}), keeping in mind that $\mbox{Det}_3$
satisfy all conditions~of (\ref{GJI}), after some trivial
evaluations we come to the following relation $(\mbox{Det}_3=\vc{1} K \ve{1})$:
\begin{equation}
\hspace*{-5pt}\begin{array}{l}
\displaystyle u(u^{-1} u_x)_y =(\mbox{Det}\, u)^{-1}
\pmatrix{ \bar d_2-\bar d_3 (X^+_1)_l, & 0 \cr
\bar d_1-\bar d_2 (X^+_1)_l , & 0 \cr}
\vspace{3mm}\\
\displaystyle \qquad \qquad \times \pmatrix{ d_2-d_3 (X^-_1)_r , & d_1 -d_2 ( X^-_1)_r \cr
0 , & 0 \cr} \vc{1} K \ve{1}.
\end{array} \label{FF}
\end{equation}
Not cumbersome transformation leads the last expression to the f\/inally form:
\begin{equation}
(u^{-1} u_x)_y =(\mbox{Det}\, u)^{-1} u^{-1}
\pmatrix{ \bar d_2, & -\bar d_3 \cr
\bar d_1, & -\bar d_2 \cr} u
\pmatrix{ d_2, & d _1 \cr
-d_3, & -d_2 \cr}. \label{B21}
\end{equation}
(\ref{B21}) is nonabelian Toda chain for $B_2$ algebra with $(0,1)$
grading. To the best of our knowledge it was not considered before.
System (\ref{B21}) is form-invariant with respect to transformation
$u\to \bar g(y) u g(x)$, with the help of which it is possible to
evaluate matrices depending on $x$, $y$ arguments to constant values. We omit
here the question about the possible canonical forms of the system
$B_2(0,1;1,1;\bar d,d)$ (\ref{B21}).
\subsection{The case of {\mathversion{bold}$G_2$} algebra}
As it is possible to expect, this case is the most cumbersome. It corresponds to
the choise $p=3$ in (\ref{RS}). Firstly, we will consider the case of
$(0,1)$ grading as the most simple one. It is connected with the 7-th
dimensional f\/irst fundamental representation of $G_2$ algebra (group).
The second one connected with $(1,0)$ grading is $14$-th dimensional.
\subsubsection{(0,1) grading}
In this case $L^{\pm}$ may contain the components $\pm 1$, $\pm 2$
graded subspaces and have the form:
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle L^+= \bar d_1 X^+_2+ \bar d_2 [X^+_1,X^+_2]+{1\over 2}\bar d_3 [X^+_1[X^+_1,
X^+_2]]
\vspace{3mm}\\
\displaystyle \qquad +{1\over 6}\bar d_4 [X^+_1[X^+_1[X^+_1,X^+_2]]]+{1\over 3} \bar d^2
[X^+_2[X^+_1[X^+_1[X^+_1,X^+_2]]]],
\end{array}
\]
$L^-=(L^+)^T$, where $T$ sign of transposition; with
simultaneously exchanging all coef\/f\/icients $\bar d\to d$. This
operation we will call as ``hermitian conjugation".
Four coef\/f\/icient functions $d_i$, $\bar d_i$ on the generators of the $\pm 1$
graded subspaces in $L^{\pm}$ are united to the ${3\over 2}$ multiplate,
with respect to gauge transformation initiated by group elements $g_0(x)$,
$\bar g_0(y)$ belonging to the f\/irst red group.
The f\/irst fundamental representation of $G_2$ algebra is $7$-th dimensional
with the basis vectors:
\[
\hspace*{-5pt}\begin{array}{l}
\ve{1}, \quad X^-_1\ve{1}, \quad X^-_2X^-_1\ve{1}, \quad X^-_1X^-_2X^-_1\ve{1}, \quad X^-_1X^-_1
X^-_2X^-_1\ve{1},
\vspace{2mm}\\
X^-_2X^-_1X^-_1X^-_2X^-_1\ve{1}, \quad X^-_1X^-_2X^-_1X^-_1X^+_2X^-_1\ve{1}.
\end{array}
\]
The action of the operators $L^{\pm}$ on $A_1$ basis of $u$ matrix is as
follows:
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle \vc{1} L^+=\vc{1} X^+_1X^+_2(\bar d_2 -\bar d_3 X^+_1+{1\over 2} \bar d_4
X^+_1X^+_1- \bar d^2 X^+_1X^+_1X^+_2),
\vspace{3mm}\\
\displaystyle \vc{1} X^+_1L^+=\vc{1} X^+_1X^+_2(\bar d_1 - \bar d_2 X^+_1+{1\over 2}
\bar d_3 X^+_1X^+_1-\bar d^2 X^+_1X^+_1X^+_2X^+_1).
\end{array}
\]
The action of the operator $L^-$ on $A_1$ basis from the left may
be obtained from the last formulae with the help of ``hermitian
conjugation'':
\[
L^-\ve{1}= (\vc{1} L^+)^T,\qquad L^-X^-_1\ve{1}= (\vc{1} X^+_1L^+)^T,\qquad
\bar d\to d.
\]
As in the previous sections the result of calculation of the main determinant
(\ref{MCC"}) it is possible to present in the form of the product of column
operator on the line one applied to the highest vector $\vc{1} K \ve{1}^2$
of the $(2,0)$ representation of $G_2$ algebra (see Appendix~II).
The line operator form is the following
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle \Bigr [d_2-d_3X^-_1+{1\over 2}d_4(X^-_1)^2-{1\over 4} d^2(2X^-_1X^-_2-3X^-_2X^-_1)
X^-_1,
\vspace{3mm}\\
\displaystyle \qquad d_1-d_2X^-_1+{1\over 2}d_3(X^-_1)^2-{1\over 4}d^2X^-_1(2X^-_1X^-_2-3X^-_2
X^-_1)X^-_1\Bigr],
\end{array}
\]
where $X^-_i\equiv (X^-_i)_r$. Nonusual (compared with the previous examples)
form of the coef\/f\/icient on $d^2$ term is the prise for $p=3$ in (\ref{RS})
in the case of $G_2$ algebra.
The column operator is obtained from the line one with the
help introduced above rules of the ``hermitian conjugation''.
For rediscovering of last symbolical expression up to the form of usual
$2\times 2$ matrix let us introduce two dimensional column vector $\bar q$
the result of the action of the column operator on the highest vector
$\vc{1} K \ve{1}^2$ divided by itself. The same in the line case will be
denoted as $q$. Explicit expressions for the line
components of $q$ have the form:
\begin{equation}\label{Q}
\hspace*{-5pt}\begin{array}{l}
\displaystyle q_1=\left(d_2+{1\over 3}d^2\alpha_{112}\right)-
2\left(d_3+{2\over 3}d^2\alpha_{12}\right)\alpha_1+\left(d_4+2\alpha_2d^2\right)\alpha_1^2,
\vspace{3mm}\\
\displaystyle q_2=\left(d_1+{1\over 3}d^2\alpha_{1112}\right)-
2\left(d_2+{1\over 3}d^2\alpha_{112}\right)\alpha_1+
\left(d_3+{2\over 3}d^2\alpha_{12}\right)\alpha_1^2
\end{array}
\end{equation}
and with the help of ``hermitian conjugation'' corresponding expressions for
the components for the column $\bar q$.
The result of the action of line operator on the highest vector in connection
with all said above is equal to the numerical line vector $\vc{1} K \ve{1}^2
(q_1, q_2)$. The action of the column operator on it may be devided on two
steps: the action on the scalar factor $\vc{1} K \ve{1}^2 $, with the f\/inally
matrix $\vc{1} K \ve{1}^2 \bar q q$ (multiplication by the law the column on
the line) and the terms with partial mutual dif\/ferentiation of the scalar and
the lines factors. All formulae for concrete calculation of such kind the
reader can f\/ind in Appendix~II. It is necessary to pay attention to the fact,
that $X^+_2 q_i=X^-_2 \bar q_i=0$, which one can check without any dif\/f\/iculties
with the help of formulae of the Appendix~I.
Gathering all these results we obtain the equation of equivalence for $u$
function:
\begin{equation}
u(u^{-1}u_x)_y={\det}^{-1}(u)\sum_{i,j,k,l} u_{ij} u_{kl} \epsilon_{ik}
\epsilon_{jl}\bar p^{ik} p^{jl}+4 d^2\bar d^2(\mbox{Det}\, (u))^{-1} u,\label{BE}
\end{equation}
where $u_{ij}$ elements of the matrix $u$, $\epsilon_{ij}$ symmetrical
tensor of the second rank with the components $\epsilon_{12}=\epsilon_{21}=-1$,
$\epsilon_{11}=\epsilon_{22}=1$, $\bar p^{ij}$, $p^{ij}$ are two-dimensional
column and line vectors correspondingly with the components (the law
of multiplication is the column on the line):
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle p^{11}=\left(d_2+{1\over 3} d^2\alpha_{112},d_1+{1\over 3} d^2\alpha_{1112}\right),
\qquad p^{22}=\left(d_4+2d^2\alpha_2,d_3+{2\over 3} d^2\alpha_{12}\right),
\vspace{3mm}\\
\displaystyle p^{12}=p^{21}=\left(d_2+{1\over 3} d^2\alpha_{112},d_3+{2\over 3} d^2\alpha_{12}\right).
\end{array}
\]
It remains only to f\/ind the derivatives $(\bar p_{ij})_x$, $(p_{kl})_y$ and convince ourselves
that together with the (\ref{BE}) they compose the closed system of equations of equivalence
or exactly integrable $G_2(0,1;2,2;\bar d^2,d^2)$ system.
Four components of $p^{22}$, $p^{11}$ with respect to transformation of the
f\/irst red algebra compose the ${3\over 2}$ spin-multiplet. So it will be
suitable to redenote them by single four-dimensional symbol $p_i$. And the
same for ``hermitian conjugating'' values $\bar p_i$.
Let us follow the calculation of $(\bar p_4)_x=2\bar d^2 (\bar \alpha_2)_x$.
The calculation of this the derivative do not dif\/ferent from the
corresponding computations of Section~3 (see (\ref{AR}) and (\ref{ARR})).
We have consequently:
\begin{equation}
(\bar \alpha_2)_x=\vc{2} K \ve{2})^{-2} \pmatrix{
\vc{2} K \ve{2}, & \vc{2} K L^- \ve{2} \cr
\vc{2} X^+_2 K \ve{2}, & \vc{2} X^+_2 K L^- \ve{2} \cr}.
\label{ARm}
\end{equation}
With the help of the technique used many times before we evaluate the last
expression to:
\[
\hspace*{-5pt}\begin{array}{l}
(\bar \alpha_2)_x=L^-_r(X^+_2)_l \ln \vc{2} K \ve{2}
=\left[d_1-d_2 X^-_1+{1\over 3} d_3 (X^-_1)^2\right.
\vspace{3mm}\\
\displaystyle \qquad \left. -{1\over 6} d_4 (X^-_1)^3+
d^2([[[X^-_2,X^-_1]X^-_1]X^-_1]-X^-_2(X^-_1)^3)\right] \theta_2.
\end{array}
\]
Using with respect to the last expression formulae of Appendix~I we come to
the system of equations of equivalence for $\bar p$ components of
${3\over 2}$ multiplet:
\begin{equation} \label{BLE}
\hspace*{-5pt}\begin{array}{l}
\displaystyle (\bar p_4)_x={2\bar d^2\over \mbox{Det}^2 (u)}\left(p_1 u_{11}^3-3p_2 u^2_{11} u_{12}+
3p_3 u_{11} u^2_{12}-p_4 u^3_{12}\right),
\vspace{3mm}\\
\displaystyle (\bar p_3)_x={2\bar d^2\over \mbox{Det}^2 (u)}\left(p_1 u_{11}^2 u_{21}-p_2 \left(u^2_{11}
u_{22}+ 2u_{11} u_{21} u_{12}\right)\right.
\vspace{3mm}\\
\displaystyle\phantom{(\bar p_2)_x=} \left. +p_3 \left(2u_{11} u_{12}u_{21}+u^2 _{12}u_{21}\right)-p_4 u^2_{12}u_{22}\right),
\vspace{3mm}\\
\displaystyle (\bar p_2)_x={2\bar d^2\over \mbox{Det}^2 (u)}
\left(p_1 u_{11} u^2_{21}-p_2 \left(u^2_{21} u_{12}+ 2u_{11} u_{21} u_{22}\right)\right.
\vspace{3mm}\\
\displaystyle \phantom{(\bar p_2)_x=} \left. +p_3 \left(2u_{22} u_{12}u_{21}+u^2 _{22}u_{11}\right)-p_4
u^2_{22}u_{12}\right),
\vspace{3mm}\\
\displaystyle (\bar p_1)_x={2\bar d^2\over \mbox{Det}^2 (u)}\left(p_1 u_{21}^3-3p_2 u^2_{21} u_{22}+
3p_3 u_{21} u^2_{22}-p_4 u^3_{22}\right).
\end{array}
\end{equation}
And corresponding system for derivatives $p_y$, which can be obtained from
(\ref{BLE}) with the help of ``hermitian conjugation''.
The symmetry of the constructed exactly integrable $G_2(0,1;2,2;\bar d^2,d^2)$
system (\ref{BE}), (\ref{BLE}) is higher than any possible espectations.
>From the physical point of view this system may be considered as the
interuction of charge ${3\over 2}$ spin particle ($p,\bar p$) with
neutral scalar-vector f\/ield $u$.
\subsubsection{(1,0) grading}
In this case $L^{\pm}$ may contain the components $\pm 1$, $\pm 2$, $\pm 3$
graded subspaces and have the form:
\[
\hspace*{-5pt}\begin{array}{l}
L^+=\bar c_1 X^+_1+ \bar c_2 [X^+_1,X^+_2]+\bar c^2 [X^+_1[X^+_1,X^+_2]]
\vspace{2mm}\\
\displaystyle \qquad + \bar c^3_1 [X^+_1[X^+_1[X^+_1,X^+_2]]]+\bar c^3_2 [X^+_2[X^+_1[X^+_1[X^+_1,
X^+_2]]]],
\end{array}
\]
$L^-=(L^+)^T$, where $T$ sign of transposition
($(X^+_i)^T=X^-_i)$; with simultaneously exchange of all
coef\/f\/icients $\bar c\to c$. This operation was called as
``hermitian conjugation'' in the previous subsection and we
conserve here this notation.
As always we begin from the equation of equivalence for two dimensional
matrix $u$ connected with the second simple root of $G_2$ algebra.
For the decoded of universal equation~(\ref{MCC"}) it is necessary
the knowledge of the action of $L^{\pm}$ on the basis. We represent below only
part of basis vectors of the second fundamental ($14$-th dimensional)
representation of $G_2$ algebra:
\[
\hspace*{-5pt}\begin{array}{l}
\ve{2}, \ X^-_2\ve{2}, \ X^-_1X^-_2\ve{2}, \ X^-_1X^1_2X^-_2\ve{2}, \
X^-_1X^-_1X^-_1X^-_2\ve{2}, \ X^-_2X^-_1X^-_1X^-_2\ve{2},
\vspace{2mm}\\
X^-_2X^-_1X^-_1X^-_1X^-_2\ve{2}, \ X^-_1X^-_2X^-_1X^-_1X^-_2\ve{2}, \
X^-_1X^-_2X^-_1X^-_1X^-_1X^-_2\ve{2}.
\end{array}
\]
The main equations (\ref{I}) are obviously invariant with the respect to
the gauge transformation iniciated by $g_0(x)$, $\bar g_0(y)$ elements of the
red algebra of the second simple root. With respect to this transformations
two coef\/f\/icients of zero $(c^1,\bar c^1)$ and third $(c^3,\bar c^3)$
order graded subspaces are transformed as spinor (anti-) multiplets; $c^2$,
$\bar c^2$ are the scalar ones. With the help of such transformation it is always
possible to satisfy the condition $c^3_2=\bar c^3_2=0$ (what is essential
simplif\/ied the calculation) and reconstruct the general case at the f\/inal
step using invariance condition.
The action of the $L^{\pm}$ operators on the basis states of the second red
algebra has the form:
\[
\hspace*{-5pt}\begin{array}{l}
\vc{2} L^+=\vc{2} X^+_2X^+_1(-\bar c^1_2 +\bar c^2 X^+_1-\bar c^3_1 X^+_1X^+_1
+\bar c^3_2 (2X^+_1X^+_1X^+_2-3X^+_1X^+_2X^+_1),
\vspace{2mm}\\
\vc{2} X^+_2L^+=\vc{2} X^+_2X^+_1(\bar c^1_1 +\bar c^2 X^+_1X^+_2+
(X^+_1X^+_1X^+_2-3X^+_1X^+_2X^+_1)(\bar c^3_1-\bar c^3_2 X^+_2).
\end{array}
\]
The action of the operator $L^-$ on $A_1$ basis from the left may be obtained
from the last formulae with the help of ``hermitian conjugation'':
\[
L^-\ve{2}= (\vc{2} L^+)^T,\qquad L^-X^-_1\ve{2}= (\vc{2} X^+_1L^+)^T,\qquad
\bar c\to c.
\]
Taking into account arguments of the Appendix~II the result of the
calculation of determinant of the third order (\ref{MCC"}) may be
presented in the operator column on line form, acting on the
highest vector of $(4,0)$ representation ($3 \vc{1} K \ve{1}^4$)
of $G_2$ algebra.
The line (``hermitian conjugating'' column) operators has the form (in this
expression we put $c^3_2=\bar c^3_2=0$):
\[
\left(-c^1_2 +c^2 X^-_1-{1\over 2}\bar c^3_1 (X^-_1)^2,\quad
c^1_1 +c^2 X^-_2X^-_1+{1\over 8}c^3_1
(X^-_2X^-_1X^-_1-6X^-_1X^-_2X^-_1)\right).
\]
Further calculations are on the level of accurate application of
dif\/ferentiation rules and combination terms of the same nature.
Equation of equivalence for $u$ function have the f\/inal form:
\begin{equation}
u(u^{-1}u_x)y=3 \, {\det}^{{1\over 3}}(u) \bar p^1 p^1+12 \, {\det}^{-{1\over 3}}(u)
\bar p^2 p^2 u+18 \,{\det}^{-1}(u) (u\bar c^3) (c^3 u),
\label{HV}
\end{equation}
where $p^1$ is the spinor with the components $p^1=(-c^1_2+4c^2\alpha_1
-6c^3_1\alpha_1^2, c^1_1+4c^2
\alpha_{21}-c^3_1(\alpha_{121}+2\alpha_1\alpha_{21}))$; scalar $p^2=
c^2-3c^3_1\alpha_1$ and corresponding expressions for bar values.
We present the system of equivalence equations without any further comments:
\begin{equation}
(\bar p^2)_x=-3 \, {\det}^{-{2\over 3}} \sum_{i,j,k,l}\bar c^3_i u_{ij}\epsilon_{kl}
p^1_l,\qquad (\bar p^1_i)_x={\det}^{-{2\over 3}}\bar p^2 \sum_{j,k,l} u_{ij}
\epsilon_{k,l}p^1_l,\label{LLL}
\end{equation}
where $\epsilon_{k,l}=-\epsilon_{l,k}$ antisymmetrical tensor of the second
rank $\epsilon_{1,2}=-\epsilon_{2,1}=1$.
And, of, course the corresponding system with the derivatives $p^1_y$, $p^2_y$.
Physical interpretation of the last system may be connected with spinor
particle interacting with charged scalar $(p^2,\bar p^2)$ and neutral
scalar-vector f\/ields in two dimensions.
\section{Concluding remarks}
In some sense in the present paper the initial idea of Sofus Lie
to introduce continuous groups as powerful apparatus for solving
the dif\/ferential equations is realized.
On the examples of semisimple groups of second order we have decoded this
idea and described explicitly exactly integrable systems whose general
solutions can be obtained with the help and in the terms of
group representation theory. We have no doubts (and partially can prove
this) that the same construction is applicable to the case of arbitrary
Lie groups and hope to prove this statement completely or to see the proof in
the literature in the nearest future.
\subsection*{Acknowledgements}
Author is indebted to the Instituto de Investigaciones en Matem'aticas
Aplicadas y en Sistemas, UNAM for beautiful conditions for his work.
Author freundly thanks N.~Ata\-ki\-shiyev for permanent discussions in the
process of working on this paper and big practical help.
This work was done under partial support of Russian Foundation of Fundamental
Researches (RFFI) GRANT.
\renewcommand{\theequation}{{\rm I}.\arabic{equation}}
\setcounter{equation}{0}
\section*{Appendix~I}
The formulae below are the general ones and have in their foundation the
f\/irst Jacobi identity only.
Let us def\/ine:
\[
\theta_j=\prod_{i=1}^r (\vc{i} G \ve{i})^{-K_{ji}}.
\]
As a result of dif\/ferentiation of $\ln \theta_i$, we obtain:
\begin{equation}
(X^-_q)_r \theta_i=-\theta_i K_{iq} \alpha_q,\qquad (X^+_q)_l \theta_i=
-\theta_i K_{iq} \bar \alpha_q, \label{AI1}
\end{equation}
\begin{equation}
(X^-_q)_r \bar \alpha_i=\delta_{q,i} \theta_i,\qquad (X^+_q)_l \alpha_i=
\delta_{q,i} \theta_i. \label{AI2}
\end{equation}
In the case of the second order algebras:
\begin{equation}
\theta_1={\vc{2} G \ve{2}\over \vc{1} G \ve{1}^2},
\qquad \theta_2={\vc{1} G \ve{1}^p\over \vc{1} G \ve{1}^2}.\label{AI3}
\end{equation}
\renewcommand{\theequation}{{\rm II}.\arabic{equation}}
\setcounter{equation}{0}
\section*{Appendix II}
Let us consider the determinant of the third order the matrix entiries
of which are coinsided with the matrix elements of $G_2$ group element $K$
taken between the bra and the ket three dimensional bases:
\begin{equation}
\vc{1},\ \vc{1} X^+_1,\ \vc{1} X^+_2X^+_1X^+_1X^+_2X^+_1,\
\ve{1}, \ X^-_1\ve{1}, \ X^-_2X^-_1X^-_1X^-_2X^-_1\ve{1}.\label{AII}
\end{equation}
Acting on such determinant by generator $(X^+_2)_r$ and taking
$(X^-_1X^-_1)_r$ out of its sign we come to the following ket basis:
\[
\ve{1}, X^-_1\ve{1}, X^-_2X^-_1\ve{1}
\]
which in connection with the (\ref{GJI}) tell us that the initial $\mbox{Det}_3$
(up to the terms anihilated by generators of the positive simple roots from
right and negative ones from the left) belongs
to $(2,0)$ ($Vh_1=2$, $Vh_2=0$) representation of $G_2$ group. For
initial determinant $Vh_1=1$, $Vh_2=0$. Each basis vector (see (\ref{high})) may be
obtained with
consequent application of the lowering operators to the higest vector
($\vc{1} K \ve{1}^2$ in the present case). There are two possibility to
combination of the lowering operators:
\[
( (AX^-_2X^-_1+BX^-_1X^-_2)X^-_1 )_r
\]
and the same expression from the left combination of the raising
generators. The condition that $\mbox{Det}_3$ is anihilated by
generators $(X^+_1)_r(X^-_1)_l$, which is a direct corollary of
the structure of the bra and ket basises, allow to f\/ind relation
between the constants $3A+2B=0$ and obtain the expression used in
the main text (\ref{Q}) and above. We obtain the following value
for $\mbox{Det}_3$ in basis (\ref{AII}):
\[
\hspace*{-5pt}\begin{array}{l}
\displaystyle \mbox{Det}_3={1\over 16}((2X^-_1X^-_2-3X^-_2X^-_1)X^-_1)
\vspace{3mm}\\
\displaystyle \phantom{\mbox{Det}_3=} \times (X^+_1(2X^+_2X^+_1- 3X^+_1X^+_2))
\vc{1} K \ve{1}^2+\vc{1} K \ve{1}.
\end{array}
\]
Below we present necessary formulae for calculation of (\ref{BE}).
We restrict ourselves by (11) component of it.
All ``mixed'' terms may be gothered in the following form:
\begin{equation}
\hspace*{-5pt}\begin{array}{l}
\displaystyle -{\bar d^2\over 4}\left[2(X^+_1X^+_2X^+_1 q_1)-3(X^+_2X^+_1X^+_1 q_1)-8\bar
\alpha_1(X^+_2X^+_1 q_1)-8\bar \alpha_1 (X^+_1 q_1)\right]
\vspace{3mm}\\
\displaystyle \qquad - \bar d_3 (X^+_1 q_1)+2\bar d_4
\bar \alpha_1(X^+_1 q_1)+{\bar d_4\over 2} ((X^+_1)^2 q_1).
\end{array}
\label{AII1}
\end{equation}
Using the def\/inition of vector $q$ (\ref{Q}) and formulae of Appendix~I,
we obtain:
\[
\hspace*{-5pt}\begin{array}{l}
(X^+_1 q_1)=2\theta_1(p^{22}_1\alpha_1-p^{22}_2)\equiv 2\theta_1 P,\quad
X^+_2 P=0,\quad (X^+_2X^+_1 q_1)=2\theta_1 \bar \alpha_2 P,
\vspace{2mm}\\
(X^+_1X^+_1 q_1)=2\theta_1^2 p^{22}_1-4\theta_1 \bar \alpha_1 P,\quad
(X^+_1X^+_2X^+_1 q_1)=2\theta_1(\bar \alpha_{21}-2 \bar \alpha_1 \bar \alpha_2
)P+2\theta_1^2 \bar \alpha_2 p^{22}_1,
\vspace{2mm}\\
(X^+_2X^+_1X^+_1 q_1)=4\theta_1^2 \bar \alpha_2 p^{22}_1+4d^2 \theta_1^2
\theta_2-4\theta_1 \bar \alpha_1 \bar \alpha_2 P-4\theta_1 \bar \alpha_{12} P.
\end{array}
\]
|
\section{Introduction}
It is well-known that crystal fields present in real materials
can influence their magnetic behavior considerably.\cite{CFbooks}
In this paper we study a ferromagnetic insulator,
with localized interacting magnetic moments, which are also
subject to a crystal field of cubic symmetry.
The finite-temperature critical
phenomena in these systems have been investigated previously within
the Landau-Ginzburg-Wilson theory,\cite{AharonyKleinert}
and the cubic universality class and its
critical exponents have been analyzed. However, in such
phenomenological
theories, the form of the order parameter is simply postulated,
using mainly symmetry arguments. To see whether such an
order really develops in the low-temperature regime one must start
with the microscopic quantum Hamiltonian.
In terms of a quantum spin Hamiltonian, the magnetic moments
are described by spin operators, and the crystal field takes the
form of a single-site term. The lowest value of $S$
for which a cubic anisotropy term has any effect is $S=2$.
Using this value our model Hamiltonian takes the form
\begin{equation}
\label{model}
H = - \sum_{\langle ij\rangle} \bbox{S}_{i} \bbox{S}_{j}
- D \sum_{i}
\left[ (S^{x}_{i})^{4} + (S^{y}_{i})^{4} +
(S^{z}_{i})^{4} \right],
\end{equation}
with $S_{i}^{\alpha}$ ($\alpha=x,y,z$) denoting
spin operators for $S=2$ at lattice site $i$.
The first term is the isotropic,
ferromagnetic exchange interaction between nearest neighbors,
and the second term represents the crystal field anisotropy;
for $S=2$ this is the most general single-site
operator with cubic symmetry. The exchange term alone produces
ferromagnetic order in the low-temperature phase; the question
we want to address is how this order is affected by the cubic
crystal field.
The above Hamiltonian has been treated within the mean-field
approximation (MFA) by many authors.\cite{SznajdValkova} It has
been found that the sign of the crystal-field constant~$D$
defines the easy axes of the magnetic ordering at low temperatures.
For $D=0$ the system is isotropic and the spontaneous
magnetization can lie along any direction;
for $D < 0$ the cube diagonals $[111]$ are preferred;
for $0 < D < (4z)/3$,
where $z$ is the number of nearest neighbors,
the directions parallel with the cube edges $[100]$
are chosen. Moreover, the MFA predicts that for
$D > (4z)/3$ the system is disordered,
irrespective of the temperature.
Recently, two of us (M.D.\ and J.S.) have investigated the above
model,\cite{OurRS}
applying the Raleyigh-Schr\"{o}dinger perturbation
theory in the limit $D \rightarrow \infty$.
This calculation is valid for large $D$,
and the following picture emerges from it. In agreement
with the MFA, for large $D$,
there is no magnetic order at any temperature,
i.e., the averages of the spin operators vanish:
\[
\langle S_{i}^{\alpha} \rangle = 0,
\hspace*{1em} \text{for} \hspace*{1em}
\alpha=x,y,z.
\]
However, in contrast to the MFA prediction,
the symmetry may still be spontaneously broken
by the appearance of a purely quadrupolar order, described by the
averages
\[
\langle (S_{i}^{x})^2 \rangle \neq
\langle (S_{i}^{y})^2 \rangle =
\langle (S_{i}^{z})^2 \rangle.
\]
(One may interchange $x$, $y$, and $z$ in the above relation.)
This purely quadrupolar phase was predicted to
exist in the model on a two-dimensional lattice at least at zero
temperature. In three dimensions\cite{OurCB}
and in higher dimensions
it should be present both in the ground state and at finite
(but low) temperatures. We suppose, that the quadrupolar phase also
exists
in one dimension at zero temperature (note that the broken symmetry
is not continuous but discrete), although in this case the
perturbation-theory arguments are weaker.
The result that there may appear a purely quadrupolar order in
the model in Eq.\ (\ref{model}), which contains only the bilinear,
``magnetic'' exchange and the single-site term, is rather
surprising. Usually, higher order exchanges are responsible for
quadrupolar ordering.\cite{SivardiereMorin}
The appearance of quadrupolar order in
the present case is entirely a quantum
effect; one would not obtain purely quadrupolar
order by simply replacing the spin operators
in Eq.\ (\ref{model}) with classical vectors,
and then seeking the configuration with minimum energy.
The result is also of methodological interest.
It is easy to understand that one cannot get purely
quadrupolar order in this model by the MFA:
only magnetization can act as a mean field when the exchange
term contains only first powers of the spin operators.
Therefore, if the purely quadrupolar phase exists,
we have the example that, in contrast with the general belief,
the MFA may give an answer which is qualitatively wrong
in high dimensions.
Unfortunately, there is no fully reliable method which would enable
us to investigate the model in two or three dimensions, for
general $D$. Therefore, the aim of the present paper is to
check and confirm the existence of the purely
quadrupolar phase numerically at least in the one-dimensional
situation at zero
temperature. We use White's density-matrix renormalization group
(DMRG) method,\cite{Whiteoriginal} which has a proven record of
extreme efficiency in dealing with such problems. The method allows
us to study the model for general $D \geq 0$, not only in the
limit of large $D$ as in Ref.\ \onlinecite{OurRS}, and we obtain
the zero temperature phase diagram, together with some properties
of
the (quantum) phase transition separating the magnetic and
quadrupolar
phases.
The paper is organized as follows: we begin with summarizing the
information obtained from the MFA and the perturbation
theory in Sec.\ \ref{sec2}. Then we discuss how long-range order
can be
detected using the DMRG in Sec.\ \ref{sec3}. Section \ref{sec4} is
devoted to the concrete numerical results, while Sec.\ V is a brief
discussion of our conclusions. We also attach an Appendix in which
numerical predictions, obtained using identical techniques, are
confronted with exact results in the test case of the 1D Ising
model in a transverse field.
\section{Predictions from MFA \protect\\
and perturbation theory}
\label{sec2}
Let us first list the eigenvalues $e_{k}$ and eigenstates
$| \psi_{k} \rangle$ of the single-site
term in the Hamiltonian in Eq.\ (\ref{model})
\begin{eqnarray}
e_{1}=-24 D & \hspace*{2.5em} & | \psi_{1} \rangle =
\frac{1}{\sqrt{2}} (| 2 \rangle + | -2 \rangle), \nonumber \\
e_{2}=-24 D & \hspace*{2.5em} & | \psi_{2} \rangle =
| 0 \rangle, \nonumber \\
e_{3}=-18 D & \hspace*{2.5em} & | \psi_{3} \rangle =
| 1 \rangle, \nonumber \\
e_{4}=-18 D & \hspace*{2.5em} & | \psi_{4} \rangle =
\frac{1}{\sqrt{2}} (| 2 \rangle - | -2 \rangle), \nonumber \\
e_{5}=-18 D & \hspace*{2.5em} & | \psi_{5} \rangle =
-| -1 \rangle,
\label{eigenstates}
\end{eqnarray}
where by $| 2 \rangle, | 1 \rangle, \ldots, | -2 \rangle$ we
denote
eigenstates of $S^{z}$. The five states are split into a doublet
and a triplet, and the doublet energy is lower for
$D > 0$. The doublet is {\em nonmagnetic}, i.e.,
\begin{equation}
\label{nonmagneticproperty}
\langle \psi_{i} | S^{\alpha} | \psi_{j} \rangle = 0,
\end{equation}
for any $i,j=1,2$ and $\alpha=x,y,z$.
Clearly, the crystal field with $D > 0$
opposes the magnetic ordering, since the nonmagnetic states
are preferred in the total wave function of the system.
For $D=0$ the model is isotropic and possesses the classical,
ferromagnetic ground state with saturated magnetization.
As mentioned in the Introduction, the MFA predicts
\cite{SznajdValkova} that the magnetic phase exists for
$0 < D < 8/3$ ($z=2$ for the chain), and that the directions of the
cube edges $[100]$ are the easy axes in this phase. According
to the MFA, the ground-state magnetization decreases with
$D$ and disappears continuously at $D=8/3$. The MFA
result that the magnetic order vanishes at some finite $D$ is
a strong argument that this is indeed the case, since the MFA
usually overestimates the tendency towards ordering. Nevertheless,
the actual
value of $D$ at which the magnetic phase ends may be much
smaller than $8/3$ (for example, in the Ising chain in a
transverse field the MFA critical field is two times larger
than the exact critical field). Whether the actual phase
transition is continuous or discontinuous is also
a matter of question. We believe that the easy axes in the
magnetic phase are given correctly by the MFA. This approximation
neglects the quantum fluctuations in the ground state, but it
seems very unlikely that these could change the easy-axis
directions,
especially in the vicinity of $D=0$, where the magnetization
is almost saturated and the quantum fluctuations are small.
Considering the large-$D$ limit now, we make use of the
perturbation theory developed in Ref.\ \onlinecite{OurRS}. Since
the two nonmagnetic states $| \psi_{1} \rangle$ and
$| \psi_{2} \rangle$ dominate for large $D$, the model
reduces to an effective two-state Hamiltonian. Restricting our
attention to the one-dimensional
situation from now on, the effective
Hamiltonian $H_{\rm eff}$ can be written in terms of Pauli
matrices $\sigma_{i}^{\alpha}$ acting on the states
$| \psi_{1} \rangle$ and $| \psi_{2} \rangle$, and has the
form\cite{OurRS}
\begin{eqnarray}
H_{\rm eff} & =
& -\tilde{J_{1}} \sum_{i}
\left( \sigma_{i}^{x} \sigma_{i+1}^{x} +
\sigma_{i}^{z} \sigma_{i+1}^{z} \right)
-\tilde{J_{2}} \sum_{i}
\left( \sigma_{i}^{y} \sigma_{i+1}^{y} \right)
\nonumber\\
&& -\tilde{J_{3}} \sum_{i}
\left( \sigma_{i}^{x} \sigma_{i+2}^{x} +
\sigma_{i}^{z} \sigma_{i+2}^{z} \right)
-\tilde{J_{4}} \sum_{i}
\left(
\sigma_{i}^{z} \sigma_{i+1}^{z} \sigma_{i+2}^{z}
\right. \nonumber\\
&& \left. \quad
- \sigma_{i}^{x} \sigma_{i+1}^{x} \sigma_{i+2}^{z}
- \sigma_{i}^{x} \sigma_{i+1}^{z} \sigma_{i+2}^{x}
- \sigma_{i}^{z} \sigma_{i+1}^{x} \sigma_{i+2}^{x}
\right)\nonumber\\
&&+ {\cal O}(\frac{1}{D^4}),
\label{Heff}
\end{eqnarray}
\narrowtext
with
\begin{eqnarray*}
\tilde{J_{1}} & = & \frac{1}{2 D} -\frac{1}{12 D^{2}}
+\frac{1}{72 D^{3}}
+ {\cal O} \left( \frac{1}{D^{4}} \right),
\\
\tilde{J_{2}} & = & \frac{1}{8 D^{2}}
+\frac{1}{32 D^{3}}
+ {\cal O} \left( \frac{1}{D^{4}} \right),
\\
\tilde{J_{3}} & = & \frac{1}{18 D^{3}}
+ {\cal O} \left( \frac{1}{D^{4}} \right),
\\
\tilde{J_{4}} & = & \frac{1}{24 D^{3}}
+ {\cal O} \left( \frac{1}{D^{4}} \right).
\end{eqnarray*}
(The exchange coupling constant $J$ from Ref.\ \onlinecite{OurRS}
is set to $1$ in the present work.)
The quadrupole operators
$[(S_{i}^{\alpha})^{2}-2]$ of the original model in Eq.\
(\ref{model})
turn out to be represented by linear combinations of Pauli matrices
as\cite{OurRS}
\begin{eqnarray}
[(S_{i}^{x})^{2}-2] \hspace*{0.8em}& \longmapsto &\hspace*{0.8em}
\sqrt{3} \sigma_{i}^{x} - \sigma_{i}^{z}
+ {\cal O}(1/ D^2), \nonumber \\
\protect
[(S_{i}^{y})^{2}-2] \hspace*{0.8em}& \longmapsto &\hspace*{0.8em}
-\sqrt{3} \sigma_{i}^{x} - \sigma_{i}^{z}
+ {\cal O}(1/ D^2), \nonumber \\
\protect
[(S_{i}^{z})^{2}-2] \hspace*{0.8em}& \longmapsto &\hspace*{0.8em}
2 \sigma_{i}^{z}
+ {\cal O}(1/ D^2).
\label{Ssigma}
\end{eqnarray}
The analogous representations of $S_{i}^{\alpha}$ and
$(S_{i}^{\alpha} S_{i}^{\beta} + S_{i}^{\beta} S_{i}^{\alpha})$,
where $ \alpha \neq \beta $, vanish up to ${\cal O}(1/D^{3})$.
In the lowest order in $1/ D$, only $\tilde{J_{1}}$
is present in the effective Hamiltonian, and $H_{\rm eff}$
describes the planar model with continuous SO(2) symmetry in the
$XZ$ plane. While this model, conventionally written as the $XY$
model, is ordered at zero temperature in two dimensions
and above, in one dimension
it is known to be critical due to enhanced quantum
fluctuations.\cite{XY,Bar-McC} The in-plane correlation function
\[
\langle \sigma_{i}^{x} \sigma_{i+l}^{x} \rangle =
\langle \sigma_{i}^{z} \sigma_{i+l}^{z} \rangle
\]
decays asymptotically as a power law $1/\sqrt{l}$, where $l$ is
the distance
between the two sites. Hence, on the basis of Eq.\ (\ref{Ssigma}),
we argue that for $D \rightarrow \infty$
a critical state, in which the quadrupolar correlation
function
\[
\left\langle \left[(S_{i}^{\alpha})^{2}-2\right]
\left[(S_{i+l}^{\alpha})^{2}-2\right] \right\rangle, \qquad
\alpha=x,y,z,
\]
decays asymptotically as $1/\sqrt{l}$, is approached.
Taking into account the higher-order terms in $H_{\rm eff}$,
we obtain the planar, $XY$-type model with small perturbations.
The three-spin term with the coupling constant $\tilde{J_{4}}$
is crucial, because it reduces the symmetry of the effective
Hamiltonian from the full rotational symmetry in the $XZ$ plane
to the symmetry of discrete rotations through angles
$\pm (2 \pi)/3$. In the one-dimensional case
the $\tilde{J_{2}}$ and $\tilde{J_{3}}$ terms are
expected to be marginally irrelevant,\cite{Nomura} but in the lack
of a
rigorous study of the off-diagonal three-point symmetry-breaking
term with $\tilde{J_{4}}$,
we can only speculate (and then eventually check numerically)
whether
the ${\cal O}(1/D^3)$ contributions are relevant or irrelevant.
If they are relevant, they can stabilize the ``magnetic
order'' (i.e., the nonvanishing averages of $\sigma_{i}^{\alpha}$)
in the effective model $H_{\rm eff}$,
since the Mermin-Wagner and Coleman's theorems
do not hold when the symmetry is not continuous but discrete.
If this is indeed the case, then Eq.\ (\ref{Ssigma}) implies
that for large $D$ a quadrupolar
order, associated with the operators $[(S_{i}^{\alpha})^{2}-2]$,
exists in the original model of Eq.\ (\ref{model}) (see
Ref.\ \onlinecite{OurRS} for details). Alternatively, if all the
perturbations turn out to be irrelevant, there should be a critical
phase for large $D$.
To summarize the above, we present the relevant (order) parameters
to be analyzed as suggested by the MFA and the perturbation theory:
\begin{eqnarray}
m^{\alpha} = \langle S^{\alpha}_{i} \rangle,
\label{defm} \qquad
q^{\alpha} = \left\langle (S^{\alpha}_{i})^{2} - 2 \right\rangle.
\label{defq}
\end{eqnarray}
These ground-state averages are
bulk values in the thermodynamic limit. Note, that in a disordered
state $m^{\alpha}=q^{\alpha}=0$ for $\alpha=x,y,z$.
Based on the above analysis we expect that for $0<D<D_m$
(where $D_m$ is probably smaller than the MFA value $8/3$) a
magnetic
phase with
\begin{equation}
\label{magneticpar}
m^{x} \neq 0, \hspace*{1em} m^{y} = m^{z} = 0,
\hspace*{1.7em} q^{x} > q^{y} = q^{z},
\end{equation}
exists, while for $D>D_m$ a purely quadrupolar phase with
\begin{equation}
\label{quadrupolarpar}
m^{x} = m^{y} = m^{z} = 0, \hspace*{1.7em}
q^{x} > q^{y} = q^{z}
\end{equation}
emerges. (The indices $x,y,z$ in the above descriptions of the
phases may be interchanged.)
In both phases the original cubic symmetry of the Hamiltonian is
spontaneously broken. Note however that the subgroup remaining
invariant is different in the two cases. As discussed above,
for large $D$ a critical (disordered) phase with
$q^{\alpha}=0$ cannot be excluded either. In any case, the values
of $q^{\alpha}$ must go to zero for $D \rightarrow \infty$.
\section{The method}
\label{sec3}
The density-matrix renormalization group (DMRG) method is one of
the
most reliable numerical techniques to study one-dimensional quantum
lattice problems. For a detailed description of the algorithm see
Ref.\ \onlinecite{Whiteoriginal}. There are different
implementations of the technique; nevertheless it seems that the
most accurate results can be
obtained by investigating finite systems with open boundary
conditions. In this setup the DMRG provides good approximations for
the ground state and low-lying excited states of long, but finite
quantum
chains. This can then be supplemented by a detailed finite size
scaling analysis using appropriate scaling assumptions.
Usually for the case of spontaneous symmetry breaking the
existence of long-range order is checked by studying
appropriate two-point correlation functions $\langle A_n
A_{n^\prime}\rangle$. The direct observation of one-point functions
$\langle A_n \rangle$ is apparently hindered
by the fact that in finite systems, where
tunneling is always finite between different vacua, the ground
state is
a symmetric combination of all possible ordering directions, and thus
naively $\langle A_n \rangle=0$. A possible way out, however, is
to break the symmetry artificially by a (small) auxiliary
field which then forces the system to make a choice between the
{\it a priori}
undetermined directions, without noticably changing the value
of the order parameter. This method has considerable advantages in the
DMRG numerics where the measurement of two-point functions is rather
awkward due to the open boundary condition.
In our implementation the symmetry-breaking auxiliary
fields are only applied to the first and last spins of the open
chain. These two {\em boundary fields} are chosen to be identical,
leaving the system symmetric for $i\to L-i$ reflections
($L$ is the chain length). Then the Hamiltonian which
is actually simulated in the numerical calculations is
\begin{equation}
H \rightarrow H - {\bf h}\cdot \left( {\bf B}_1 +{\bf
B}_L\right)
\label{bf-setup}
\end{equation}
where ${\bf B}_i$ and ${\bf h}$ are vectors of appropriate
single-site operators and boundary fields. Since the system is not
translation invariant, the one-point function $\langle A_n \rangle$
depends on the position $n$ near
the chain ends, but approaches its bulk value in the middle of the
system. If the system size is much larger than the correlation
length
the one-point function rapidly saturates and forms a plateau.
Alternatively, if the chain length is too small comparing to the
correlation length (e.g., the system is critical or it is close to
criticality) no plateau appears to develop.
There are two ways to estimate the order parameter and the
associated correlation length from one-point functions. One can
define the value $a_{L/2}\equiv \langle A_{L/2} \rangle$
measured in the middle of the
chain and study its dependence on the total chain length $L$.
It is expected that $a_{L/2}(L)\to a$ (with $a$ the bulk value
of the order
parameter) exponentially fast as $L\to\infty$, whenever the system
is away from criticality, while the convergence is only algebraic
at
criticality. A disadvantage of this method is that many independent
DMRG runs must be done with different $L$ values, or one is forced
to use the ``infinite lattice'' algorithm which is less precise.
Alternatively, one can analyze the {\em profiles}
$a_n\equiv \langle A_n \rangle$ as a function of
$n$ at a fixed length $L$, providing that $L$ exceeds the
correlation
length $\xi$ reasonably (which can be checked {\it a posteriori}).
In this latter case the profile is expected to have the following
asymptotic form far from the chain ends
\begin{equation}
\label{fittingformula}
a_n = a + c \frac{e^{-n / \xi}}{n^\chi} +
c \frac{e^{-(L+1-n) / \xi}}{(L+1-n)^{\chi}} ,
\end{equation}
where $a$ is the bulk value of the order parameter, $\xi$ is the
associated correlation length, $\chi$ is a suitable exponent, and
$c$ is a constant. In general one can assume that the
order parameter and the correlation length measured this way
are identical to their bulk values
derived from two-point correlation functions. However, the exponent
of the polynomial prefactor $\chi$ is normally different from
its bulk value and may depend on the boundary conditions chosen.
In practice we analyze our numerical data by dividing the chain
into smaller segments with a length
of the order of $\xi$, and perform local least-squares fits using
the formula
in Eq.\ (\ref{fittingformula}). We have found it appropriate
to fix $\chi$ first and then look for its value which makes the
other
fitting parameters the most stable in different segments as
$n\to\infty$.
Our procedure of determining order parameters and correlation
lengths was extensively tested on a simpler problem, the 1D Ising
model in a transverse field (ITF), where exact answers to many
questions of interest are available. We summarize our experience
with this model in the Appendix.
In most cases the DMRG is very efficient to calculate spectral
gaps. The method is less useful, however, if the ground state is
(asymptotically) highly degenerate, since then several states
have to be targeted together,
resulting in a rapid drop in accuracy.
In the cubic model under investigation we expect a 6-fold
ground-state degeneracy in the magnetic phase,
and a 3-fold degeneracy
in the quadrupolar phase, in accordance with the number of
easy-axis directions
for the two order parameters. The
6-fold degenerate case already exceeded our computational
limitations, thus some of our conclusions with respect to
that case are based on
short chain exact diagonalization results only.
Errors in the quantitative predictions we report in the next
sections stem from two sources, either from the limited accuracy
of the numerical technique we
have been using, or from the approximate (asymptotic) nature of the
formulas we applied for fitting and extrapolating the data. As for
the DMRG errors, our results are somewhat more precise when the
``finite
lattice'' algorithm was used. In this case several
iterations were done until convergence was reached.
In principle, results obtained through the ``infinite
lattice'' method inevitably contain a systematic ``environment''
error, and thus should be treated with less confidence. However,
in the present problem improvements due to the ``finite
lattice'' iterations turned out to be rather small,
providing enough ground to assume that even our ``infinite
lattice''
results have sufficient accuracy.
In general we made various runs keeping different
number of states $M$, and extrapolated our results for
$M\to\infty$, or at least checked that the results had been
converged in
this parameter. Typically, we found good convergence in $M$
despite the fact that the maximum number of states kept did not
exceed $M=85$. Note that the number of degrees of freedom
at a single site is five in the $S=2$ case, and the model in
question has no continuous axial symmetry
(total $S^z$ is not conserved)
which could have been utilized to facilitate the DMRG calculation
in the usual way.
\section{Numerical results}
\label{sec4}
In order to obtain the zero temperature phase diagram and critical
properties of the model we carried out extensive numerical
calculations
using exact diagonalization techniques on short chains and the DMRG
method on longer systems. In the case of open boundary conditions
with
boundary fields the actual form of the Hamiltonian used in the
simulations was
\begin{eqnarray}
\label{actmodel}
H &=& - \sum_{i=1}^{L-1} \bbox{S}_{i} \bbox{S}_{i+1}
- \sum_{i=1}^L D_i
\left[ (S^{x}_{i})^{4} + (S^{y}_{i})^{4} +
(S^{z}_{i})^{4} \right] \nonumber\\
&&-{\bf h}\cdot {\bf (B_1+B_L)}
\end{eqnarray}
with a reduced cubic crystal field at the chain ends (to
counterbalance the missing bonds for the first and the last spins)
\begin{equation}
D_i = \left\{ \begin{array}{ll} D & {\rm if}\; i=2,\dots,L-1 \\
D/2 & {\rm if}\; i=1,L \end{array}
\right.
\end{equation}
and a general symmetry-breaking boundary field containing magnetic
and quadrupolar terms
\begin{equation}
{\bf B} = \left( \begin{array}{c}
S^x\\S^z\\(S^x)^2\\(S^y)^2\\(S^z)^2 \end{array} \right),
\qquad
{\bf h} = \left( \begin{array}{c}
h_x\\h_z\\h_{xx}\\h_{yy}\\h_{zz} \end{array} \right).
\label{Bh}
\end{equation}
In order to avoid handling complex numbers, we did not include a
boundary term proportional to $S^y$.
After the ground state had been found, we measured the
order parameters. To facilitate further discussion we introduce the
notations
\begin{eqnarray}
m_{n}^{\alpha} = \langle S_{n}^{\alpha} \rangle, \qquad
q_{n}^{\alpha} = \langle (S_{n}^{\alpha})^{2}-2 \rangle
\label{defmq}
\end{eqnarray}
with $n=1,\dots,L$; $\alpha=x,y,z$.
\subsection{Ground state degeneracies and gaps}
Figure \ref{fig:degen}(a) shows the lowest energy levels in a
periodic chain with $L=8$ sites obtained using
exact diagonalization techniques. Although the chain is short, the
6-fold degeneracy, consisted of a singlet, a doublet, and a triplet
is clearly discernible for small values of $D$.
In the large $D$ (quadrupolar) region the degeneracy seems
to be 3-fold, consisted of the singlet and the doublet only. In
both regimes we observe signs of a gap above these states, although
the number of data points ($L=4,6,8$) available for a given $D$
did not allow us to perform a detailed finite-size scaling study.
Note that since the
broken symmetry is discrete we do not expect any massless Goldstone
modes in the ordered regimes.
Excited states are rather massive domain walls, which
separate domains with different ordering directions, and
propagate in the system.
\begin{figure}[hbt]
\epsfxsize=\columnwidth\epsfbox{fig1b.eps}
\caption{(a) Low-energy spectrum vs cubic anisotropy $D$ in a short
chain with $L=8$ using periodic boundary condition. Shading
schematically indicates higher-energy states.
(b) Ground state--first excited state (singlet--doublet) gap (1st
XS),
and ground state--second excited state
gap (2nd XS) as a function of $1/L$ (solid lines)
and $1/L^2$ (dashed lines) at $D=1.3$.
The ticks in the horizontal axes correspond to $L=10,20,\ldots,50$.
\label{fig:degen} }
\end{figure}
As was discussed in the former section, we could only apply the
DMRG for the 3-fold degenerate case. We calculated the low-lying
four states in an open chain
at $D=1.3$, which is above the critical point $D_m$ (see later
sections), by targeting four states without any boundary fields.
The results, presented in Fig.\ \ref{fig:degen}(b), strongly
support the 3-fold asymptotic degeneracy with
a finite gap $\Delta\approx 0.08\pm 0.02$ above it in
the thermodynamic limit.
Unfortunately, this calculation was rather time consuming,
which impeded us to repeat it for other $D$ values.
\subsection{Magnetic order}
\label{sec4B}
Let us now investigate the spontaneous magnetization and the
corresponding correlation length. In these calculations, purely
magnetic boundary fields were used, i.e.,
$h_{xx}=h_{yy}=h_{zz}=0$ in Eq.\ (\ref{Bh}).
\begin{figure}[hbt]
\epsfxsize=\columnwidth\epsfbox{fig2b.eps}
\caption{(a) The parameter $m^{x}_{L/2}$ vs the chain length $L$,
and (b) the profile $m^{x}_n$ vs the position $n$ at $L$
fixed, for some representative values of $D$. Purely
magnetic boundary fields applied along the $x$ axis as given by
$h_{x}$.
\label{fig:magn} }
\end{figure}
Expecting the magnetic easy axes along the cube edges we started
with applying a boundary field along the $x$ axis setting
$h_{x}=10$ or $0.1$, $h_{z}=0$. Due to the symmetry,
with such a boundary field one must obtain $m_{n}^{x} \neq 0$
and $m_{n}^{y}=m_{n}^{z}=0$.
Figure \ref{fig:magn}(a) shows how the parameter $m_{L/2}^{x}$,
measured in the center of the chain, changes with the total
chain length $L$ for representative values of the crystal-field
strength $D$. One can see that for $D=0.5,1,1.2$ the parameter
$m_{L/2}^{x}$ converges
to finite values independent of $h_{x}$. This clearly indicates
a finite magnetic order for these $D$ values. On the other hand,
for $D=1.3,1.5,2$ the parameter $m_{L/2}^{x}$ goes to zero as
$L\to\infty$, meaning no magnetic order.
As characteristic examples the magnetization profiles
$m_{n}^{x}$ with $L$ fixed are shown in
Fig.\ \ref{fig:magn}(b) for $D=1$ and $2$.
For $D=1$ a plateau corresponding to a finite spontaneous
magnetization $m^{x}\approx 1.76$, while for $D=2$ a plateau
with $m^{x}=0$ can be observed.
So far we have assumed that the cube edges give the easy axes in
the magnetic phase. To verify this
assumption, we also applied boundary fields which do not
coincide with the expected easy directions. Choosing $h_{x}=0.04$
and $h_{z}=0.03$ we found that for $L\to\infty$ the parameter
$m_{L/2}^{x}$ always tends to the same value
as found above, while $m_{L/2}^{z}$ goes to zero, in agreement with
Eq.\ (\ref{magneticpar}).
\begin{figure}[hbt]
\epsfxsize=\columnwidth\epsfbox{fig3b.eps}
\caption{The spontaneous magnetization $m$ (left axis, dots) and
the corresponding correlation length $\xi_{m}$ (right axis,
triangles) as functions of $D$.
Notice the large error bars for $\xi_{m}$ in the magnetic phase.
\label{fig:magnD} }
\end{figure}
The above calculations allowed us to determine how the bulk
spontaneous magnetization $m$ changes with the crystal-field
strength $D$. We assume that monotonic dependencies of
$m^{\alpha}_{L/2}$ on $L$, like those in Fig.\ \ref{fig:magn}(a),
provide correct bounds for $m$, as was found for the ITF chain
in the Appendix. The magnetization $m$ is plotted in
Fig.\ \ref{fig:magnD} as a function of
$D$, together with the corresponding correlation length $\xi_m$,
discussed in details below. Since $\xi_m$ increases around the
phase transition, very long chains have to be studied in order
to find the large-$L$ limit of $m_{L/2}^{x}$ in this region.
The largest $D$
for which we can see the magnetic order is $D=1.23$. For this
$D$, taking $L=400$, we obtain a finite $m$ with an uncertainty
of $0.003$ (for smaller $D$, shorter chains are sufficient to
calculate $m$ much more precisely). Due to the increasing
correlation length, and thus computational limitations,
we were not able to trace the dependence of
$m$ on $D$ further. Starting from $D=1.26$, the phase with no
magnetic order is observed.
Although it cannot be shown directly that the spontaneous
magnetization disappears continuously at the phase-transition
point, the sharp increase of the correlation length suggests
that the transition is continuous. We therefore attempt to fit
the dependence of $m$ on $D$, shown in Fig.\ \ref{fig:magnD},
assuming a power
law singularity and a linear regular term
\begin{equation}
\label{fitm}
m(D) = c_1 (D_m-D)^{\beta} + c_2 (D_m-D).
\end{equation}
The fit is very good and we obtain
\begin{eqnarray*}
D_m = 1.2374 \pm 0.0004,\qquad \beta = 0.127 \pm 0.004,
\end{eqnarray*}
$c_1 = 2.15 \pm 0.05$, $c_2 = -0.17 \pm 0.05$.
Interestingly, the exponent $\beta$ is very
close to that of the two-dimensional Ising model. The obtained
parameter $D_m$ is our best estimate of the phase-transition
point.
To calculate the magnetic correlation length $\xi_{m}$, we analyze
the profiles of $m_{n}^{x}$ at fixed $L$, analogous to those
presented in Fig.\ \ref{fig:magn}(b). As was described earlier, we
divide the chain into short sections and perform local fits
using the formula in Eq.\ (\ref{fittingformula}).
For each $D>D_m$, i.e., in the phase with no magnetic order,
the local values of the fitting parameters obtained with
$\chi=0$ are very stable, and the convergence with increasing $n$
(towards the center of the chain) is convincing. The fitted order
parameter $m$ is always very close to zero.
It seems that the decay of
$m_{n}^{x}$ is asymptotically purely exponential, like the decay
of $m^{\rm ITF}_n$ in the disordered phase of the ITF chain,
discussed in the Appendix.
Due to the stability of local $\xi$'s, the correlation length
$\xi_{m}$ in the phase with no magnetic order can be calculated
very precisely (at $D=1.27$ the error in $\xi_{m}$ is
around 0.1\% and it is smaller for larger $D$).
In the magnetically ordered phase, for $D<D_m$, the local
fitting parameters obtained with $\chi=0$ are not stable. The
local $\xi$'s increase towards the center of the chain, like in
the ordered phase of the ITF chain, suggesting the existence of
an algebraic prefactor. To calculate $\xi_m$, we vary $\chi$ and
look for its value where the local $\xi$'s are most stable. We
do not find a universal prefactor, and the obtained $\chi$
depends on $D$ and $h_x$. However, the estimates of $\xi_m$ are
independent of the boundary field, which justifies the procedure;
the uncertainties in $\xi_m$ are around $10$\%. (The same procedure
was tested for the ordered phase of the ITF chain and yielded
acceptable estimates of the correlation length --- see Appendix.)
The dependence of the correlation length $\xi_{m}$ on the
crystal-field strength $D$ is plotted in Fig.\ \ref{fig:magnD}.
Due to the large error bars for $D<D_m$, the
functional dependence can only be analyzed
for $D>D_m$. Here, assuming a second order transition again, we fit
the dependence of $\xi_{m}$ on $D$ with a power law
\begin{equation}
\label{fitxim}
\xi_{m} (D) = c_3 (D-D_m')^{-\nu}.
\end{equation}
The fit is again rather good, yielding the values
\begin{eqnarray*}
D_m' = 1.236 \pm 0.004, \qquad \nu = 1.02 \pm 0.06.
\end{eqnarray*}
Thus, close to the phase transition both the spontaneous
magnetization and the corresponding correlation length
seem to be well described by power laws. The two independent
estimates of the phase-transition point, $D_m$ and $D_m'$
from Eqs.\ (\ref{fitm}) and (\ref{fitxim}), nicely coincide
with each other. (In what follows, we will use $D_m$, which
is more precise.) Moreover,
the exponents $\beta$ and $\nu$ are close to those of the
two-dimensional Ising model. Altogether, we have found strong
indications that, as far as the magnetic quantities $m$ and
$\xi_{m}$
are concerned, the phase transition is continuous and
is described by the critical exponents of the two-dimensional
Ising model.
\subsection{Quadrupolar order}
\label{sec4C}
We now investigate the
quadrupolar order, associated with the parameter $q^{\alpha}$
defined in Eq.\ (\ref{defq}). While the quadrupolar order must be
present in the magnetic phase ($D<D_m$), the question whether
it exists in the phase with no magnetic order ($D>D_m$) is
of central interest.
Expecting the quadrupolar easy axes along the cube edges,
as given in Eqs.\ (\ref{magneticpar}) and (\ref{quadrupolarpar}),
we began with applying boundary fields along the $x$ axis. These
were purely quadrupolar:
$h_{\alpha}=0$; $h_{xx} \neq 0$, $h_{yy}=h_{zz}=0$ in Eq.\
(\ref{Bh}); or purely magnetic:
$h_{x} \neq 0$, $h_{z}=0$; $h_{\alpha \alpha}=0$. With such
boundary fields, due to the symmetry, one must obtain
$q_n^x \neq q_n^y=q_n^z$.
\begin{figure}[hbt]
\epsfxsize=\columnwidth\epsfbox{fig4b.eps}
\caption{(a) The parameter $q^{x}_{L/2}$ vs the chain length $L$,
and (b) the profile $q^{x}_n$ vs the position $n$ at
$L$ fixed, for two values of $D$.
Purely magnetic ($h_{x}$) and purely quadrupolar ($h_{xx}$)
boundary fields applied along the $x$ axis.
\label{fig:quad} }
\end{figure}
We found that
in both phases the order parameter $q_{L/2}^{x}$ measured in
the chain center converges with increasing $L$ to finite
values which do not depend on the actual boundary fields.
Figure \ref{fig:quad}(a) shows
these dependencies for $D=1.2<D_m$ and $D=1.3>D_m$.
The appropriate profiles of $q^{x}_{n}$ with
$L$ fixed are plotted in Fig.\ \ref{fig:quad}(b).
In both phases the picture is
typical for an ordered state: $q^{x}_{n}$ has plateaus
in the middle of the chain, which correspond to finite bulk
$q^{\alpha}$. Thus, we find quadrupolar order on both sides of
$D_m$,
confirming the existence of a purely quadrupolar phase for $D>D_m$.
To verify that the cube edges give the easy axes of the quadrupolar
ordering, we applied a
quadrupolar boundary field having all three components
different from zero: $h_{\alpha}=0$, $h_{xx}=0.03$, $h_{yy}=0.02$,
and $h_{zz}=0.01$. Then, in both phases, it was observed that
$q^{x}_{L/2}>q^{y}_{L/2}=q^{z}_{L/2}$ in the large-$L$ limit,
and the limiting values of $q^{\alpha}_{L/2}$ were the same as
found above, in accordance with Eqs. (\ref{magneticpar})
and (\ref{quadrupolarpar}).
Assuming that monotonic dependencies of $q^{\alpha}_{L/2}$ on $L$
provide correct bounds for the bulk value $q$, we determined
$q$ for various values of $D$.
($q$ is defined as the largest of $q^{\alpha}$ in a
broken-symmetry state.) The results are
depicted as bold points in Fig.\ \ref{fig:quadD}; the precision
in $q$ is $10^{-3}$ or better. We observed that,
starting from $D \approx 1.3$, the associated correlation length
$\xi_q$ increases rapidly with $D$
(see the inset in Fig.\ \ref{fig:quadD}).
At the same time,
the DMRG precision for $q^{\alpha}_{n}$ fell dramatically.
For these reasons, already at $D=1.45$ (and for larger $D$ too)
we were unable to obtain precise results
for a chain long enough to show us the large-$L$ limit of
$q^{\alpha}_{L/2}$ (for $D=1.45$ and $1.5$ we obtained upper bounds
on $q$, which are shown as open circles in Fig.\ \ref{fig:quadD}).
We had similar problems close to the transition point $D_m$.
\begin{figure}[hbt]
\epsfxsize=\columnwidth\epsfbox{fig5b.eps}
\caption{The quadrupolar order parameter $q$ as a function of $D$.
For $D=1.45$ and $1.5$ upper bounds on $q$ are depicted as the
open circles. The inset shows very rough estimates of the
corresponding correlation length $\xi_{q}$.
\label{fig:quadD} }
\end{figure}
The effective Hamiltonian $H_{\rm eff}$ in Eq.\ (\ref{Heff})
may help to understand the difficulties for $D\agt 1.4$. Notice,
that the symmetry-breaking term $\tilde{J_4}$ in $H_{\rm eff}$ is
very small, $\tilde{J_4}/\tilde{J_1}={\cal O}(1/D^2)$.
Thus even if $\tilde{J_4}$ is relevant, the correlation
length is expected to diverge rather rapidly as $D\to\infty$.
On the other hand, the observed abrupt increase in $\xi_{q}$
and the emerging
numerical difficulties beyond $D\sim 1.4$ may indicate an
additional phase transition in which the quadrupolar long-range
order vanishes, and gives rise to an extended critical phase
with algebraically decaying quadrupolar correlations
beyond a critical
value $D_q$. A naive linear fit to the data points for $q$ in Fig.\
\ref{fig:quadD} would yield a value $D_q\approx 2.1$, but we are
not in a position to give any firm claims in this question.
Note however that if $D_q=\infty$,
i.e., if the quadrupolar phase extends
to $D\to\infty$, the curve $q(D)$ should finally bend upwards, and
in the data points in Fig.\ \ref{fig:quadD},
together with the upper bounds at $D=1.45$ and $1.5$,
the (small) curvature is consistently downwards.
Now we analyze the dependence of $q$ on $D$ in Fig.\
\ref{fig:quadD}
close to the phase-transition point $D_m$.
For $D<D_m$ we assume a power-law singularity and a linear
regular term
\begin{equation}
\label{fitqleft}
q(D) = q_{\rm left} + c_4 (D_m-D)^{\theta} + c_5 (D_m-D).
\end{equation}
Fitting yields
\begin{eqnarray*}
q_{\rm left} = 1.838 \pm 0.01, \qquad \theta = 0.5 \pm 0.1,
\end{eqnarray*}
$c_4 = 0.44 \pm 0.1$, $c_5 = -0.36 \pm 0.1$.
Thus, on this side of the phase-transition point, the quadrupolar
order parameter appears to have singular behavior, but the exponent
$\theta$ differs from $\beta$ describing the spontaneous
magnetization.
For $D>D_m$, we suppose the behavior is regular, which is
suggested by Fig.\ \ref{fig:quadD}. Our fitting formula is now
\begin{equation}
\label{fitqright}
q(D) = q_{\rm right} + c_6 (D-D_m) + c_7 (D-D_m)^{2},
\end{equation}
giving
\begin{eqnarray*}
q_{\rm right} & = & 1.840 \pm 0.002,
\end{eqnarray*}
$c_6 = -1.83 \pm 0.03$, $c_7 = -0.6 \pm 0.1$.
The left and right estimates of the order parameter
$q$ at $D_m$ are in agreement, which supports our scaling
assumptions.
However, we cannot exclude the possibility that a singular
derivative
in $q$ exists also for $D>D_m$, so close to $D_m$ that we
do not observe it in our numerics.
To calculate the correlation length $\xi_{q}$, we analyze
the profiles of $q^{\alpha}_{n}$, like those in Fig.\
\ref{fig:quad}(b). In our opinion,
only a purely quadrupolar boundary field is suitable for this
purpose. With a magnetic boundary field, the decay of
$q_{n}^{\alpha}$ could be driven by the decay of $m_{n}^{\alpha}$,
and so governed by $\xi_{m}$ and not $\xi_{q}$.
Using the formula in Eq.\ (\ref{fittingformula}) again, we find
that
both for $D<D_m$ and $D>D_m$ the prefactor exponent $\chi$ is
nonzero and it depends on $D$ and the boundary field. The fitting
yields rough estimates to $\xi_q$ (Fig.\ \ref{fig:quadD} inset),
the errors are believed to be around $20$\%.
The results suggest that $\xi_{m} = \xi_{q}$ for
$D<D_m$, which is the expected answer, since the magnetic and
quadrupolar order is intimately connected in this phase.
For $D>D_m$, the correlation length $\xi_q$ increases with $D$
starting from $D\approx 1.3$, as mentioned above, but, due to the
poor precision, we cannot resolve whether $\xi_{q}$ diverges
(numerically it increases sharply) for $D \rightarrow D_m$
or it remains finite.
\section{Conclusions}
In summary, we have studied the ground-state properties of a
one-dimensional $S=2$ ferromagnetic spin chain with single-site
cubic crystal-field anisotropy $D$.
We argued that in contrast with the mean-field
prediction, perturbation theory suggests
the possibility that a purely quadrupolar phase exists for
large values of $D$. This quadrupolar phase was expected to be
separated
from the magnetic phase by a continuous quantum phase transition at
a critical point $D_m$.
We verified this conjecture by investigating the model
numerically using the density-matrix renormalization group method
on open chains with special, symmetry-breaking boundary conditions.
In most cases, the method provided precise estimates of the
magnetic and quadrupolar order parameters. Very close to the
phase-transition point and in the large-$D$ limit, however,
our results were
less accurate due to the rapidly increasing correlation lengths.
For the correlation lengths themselves we could only obtain rather
rough estimates.
Evidence has been obtained that, in qualitative agreement with the
mean-field prediction, the spontaneous magnetization diminishes
continuously at the phase-transition point. Regarding the magnetic
properties, the transition seems to be characterized by the
critical exponents of the two-dimensional Ising model.
This could be plausible, since the extra
broken symmetry of the magnetic ground state with respect
to that of the quadrupolar ground state is just a $Z_2$
subgroup --- the same group as the one breaking down spontaneously
in the case of the Ising model.
Both in the magnetic and nonmagnetic phases
we demonstrated the presence of a quadrupolar order.
In the former case the quadrupolar order is just a
``secondary'' effect,
inevitably present in any $S\ge 1$ model with magnetic order.
In the latter case, however, it is the ``primary'' order,
constituting a qualitatively different broken-symmetry phase.
At the phase-transition point $D_m$, the quadrupolar order
parameter is continuous. For $D\alt D_m$, we could clearly
discern a singularity in the derivative of the order parameter,
and found that the quadrupolar correlation length diverges together
with the magnetic correlation length.
For $D\agt D_m$, our observations are much less concrete: the
order parameter can be fitted reasonably well by a low-order
(regular) polynomial without any singular terms, and,
although the quadrupolar correlation length
increases as $D_m$ is approached from above, our results
do not suggest unambiguously a divergence on
this side. It is not clear for us whether the phase transition
involving the magnetization should have any precursor in the
quadrupolar fluctuations above $D_m$.
Mainly due to computational limitations we were unable to resolve
convincingly the question whether the quadrupolar phase extends
to $D\to \infty$ or there is a finite value $D_q$ where
the quadrupolar
long-range order disappears. This question is unique to the
one-dimensional situation --- in two dimensions
and higher we expect a finite
quadrupolar order at zero temperature for any $D>D_m$.
Finally, the scenario for the magnetic-to-quadrupolar phase
transition
was supported by the analysis of the degeneracy of the ground state
as a function of $D$. We observed a 6-fold and a 3-fold asymptotic
degeneracy below and above $D_m$ respectively,
in accordance with the
the number of possible ordering directions in the two phases.
The confirmation of the purely quadrupolar phase in the
one-dimensional
case gives rise to the belief that such a phase
should also exist in higher dimensions at sufficiently low
temperatures
when the cubic crystal field is strong. Increasing temperature
necessarily destroys the quadrupolar order and leads to a
finite-temperature phase transition between the quadrupolar
and a completely disordered phase. Note that the
correct critical theory of this
transition cannot be obtained by simply substituting spins with
classical vectors in our model,
since this approach is unable to account for the
purely quadrupolar order.
The usual practice of neglecting quantum fluctuations by treating
spins classically around a finite-temperature phase transition
would confront fundamental difficulties in this case.
\acknowledgements
The authors thank Steven White for the suggestion of applying
symmetry-breaking boundary conditions. Valuable discussions with
Enrico Carlon, Andrzej Drzewi\'{n}ski, and Uli Schollw\"{o}ck
are also acknowledged. This research was supported by
the KBN (Poland) grant no.\ 2P03B07214
and the EPSRC (UK) grant no.\ GR/L55346.
\begin{appendix}
\section*{A test calculation}
Our numerical procedure and extrapolation methods were tested on
the zero-temperature 1D Ising model in a transverse field (ITF)
where an exact solution for the bulk quantities
is available.\cite{ITF,Bar-McC} Our experience with this model
provided useful guidelines to refine our
procedure in the study of the cubic model of Eq.\ (\ref{model}).
The Hamiltonian of the ITF model on an open chain with length $L$
is defined as
\begin{eqnarray}
\label{HITF}
H^{\rm ITF} &=& - \sum_{i=1}^{L-1} \sigma_{i}^{z}
\sigma_{i+1}^{z}
- \sum_{i=1}^L \Gamma_i
\sigma_{i}^{x}\\
\Gamma_i &=& \left\{ \begin{array}{ll} \Gamma & {\rm if}\;
i=2,\dots,L-1 \\
\Gamma/2 & {\rm if}\; i=1,L
\end{array}
\right. \nonumber
\end{eqnarray}
where $\sigma_{i}^{\alpha}$ denotes Pauli matrices at site $i$,
and $\Gamma$ is the transverse magnetic field. Note that in the
above
definition the spins at the chain ends, being coupled to one
neighbor only,
are subject to the reduced field $\Gamma/2$.
We concentrate on the spontaneous magnetization
$ m^{\rm ITF} = \langle \sigma_{i}^{z} \rangle $
and the corresponding correlation length $\xi^{\rm ITF}$.
For $\Gamma=0$ the Hamiltonian $H^{\rm ITF}$ has the classical,
ferromagnetic ground state with $m^{\rm ITF} = \pm 1$, and
the single-site term in $H^{\rm ITF}$ tends
to destroy the spontaneous magnetization; in this respect the
ITF model is similar to the cubic model in Eq.\ (\ref{model}).
In the thermodynamic limit there are two phases in the ITF chain,
distinguished by the order parameter $m^{\rm ITF}$. For
$0 < \Gamma < 1$ the ground state is ordered due to spontaneous
symmetry
breaking
\begin{equation}
m^{\rm ITF} = \pm \left(1-\Gamma^{2}\right)^{1/8},
\hspace*{2em}
\xi^{\rm ITF} = - \frac{1}{2 \log \Gamma},
\label{orderedITF}
\end{equation}
while for $\Gamma>1$ there is a disordered phase with
\begin{equation}
m^{\rm ITF} = 0,
\hspace*{2em}
\xi^{\rm ITF} = \frac{1}{\log \Gamma}.
\label{disorderedITF}
\end{equation}
At $\Gamma=1$ the ground state is critical and $\xi^{\rm ITF} =
\infty$.
In our study the following values of the transverse field were
considered:
$\Gamma=0.9$ (ordered phase) where $\xi^{\rm ITF} \approx 4.75$;
$\Gamma=1.1$ (disordered phase) where $\xi^{\rm ITF} \approx
10.49$;
and the critical point $\Gamma=1$ with $\xi^{\rm ITF}=\infty$.
We computed the position dependent order parameter
$m^{\rm ITF}_{n} = \langle \sigma_{n}^{z} \rangle$
using the DMRG method with boundary fields up to $L=200$.
As a symmetry-breaking boundary field we considered
\begin{equation}
B = \sigma^z
\end{equation}
in Eq.\ (\ref{bf-setup}), and we used two different values of $h$:
a strong $h=10$ and a weak $h=0.1$ boundary field.
In the case of the ITF chain
the DMRG is extremely precise even close to
the critical point.\cite{Gaborandcousin} Keeping a relatively small
number of states $M=64$ the
truncation errors are already negligible and the fitting procedure
can be tested on numerically exact data. We emphasize that although
rigorous results are available for quantities in
the thermodynamic limit, the
behavior of relatively short chains with complicated boundary
conditions is not feasible for a study with purely
analytical tools even for the ITF model.
\begin{figure}[hbt]
\epsfxsize=\columnwidth\epsfbox{fig6b.eps}
\caption{ITF chain results. (a) Order parameter $m^{\rm
ITF}_{L/2}$ measured in the center of the chain vs the chain length
$L$. (b) Order parameter profiles $m^{\rm ITF}_{n}$ as a function
of position in the chain $n$ at fixed $L=200$.
Curves are labeled by the values of the
transverse field $\Gamma$ and the boundary-field strength $h$.
\label{fig:ITF} }
\end{figure}
The value of the order parameter at one of the central spins
$m^{\rm ITF}_{L/2}$ as a function of the chain length $L$ is
shown in Fig.\ \ref{fig:ITF}(a). For these data the
``infinite-lattice''
DMRG was used from $L=4$ to $L=200$, thus a systematic
``environment error'' cannot be excluded.\cite{Gaborandcousin}
In the ordered phase, $m^{\rm ITF}_{L/2}$ converges to the
same finite value for both values of $h$, and the two dependencies
are monotonic. At each $L$, correct bounds for the bulk
$m^{\rm ITF}$ in Eq.\ (\ref{orderedITF}) are obtained,
and at $L=200$, ten digits of $m^{\rm ITF}$ are recovered.
In the disordered phase, $m^{\rm ITF}_{L/2}$ converges
to zero for both boundary fields $h$. At the critical point,
the limit of $m^{\rm ITF}_{L/2}$ does not
show up in Fig.\ \ref{fig:ITF}(a), we can only say that the
correlation length must be of the order of $L=200$ or larger.
In Fig.\ \ref{fig:ITF}(b) we show the order parameter profiles
$m^{\rm ITF}_{n}$ for a chain length fixed at $L=200$. This length
exceeds the exact correlation lengths $\xi^{\rm ITF}$ for
$\Gamma=0.9$
and $\Gamma=1.1$ many times.
There are plateaus for $\Gamma=0.9$ and $\Gamma=1.1$,
which correspond to the bulk values of the order parameter.
There is no plateau at the critical point where the correlation
length
is infinite and the dependence of the profile is algebraic.
In order to calculate the correlation length $\xi^{\rm ITF}$,
we analyze the profiles from Fig.\ \ref{fig:ITF}(b) using the
ansatz in
Eq.\ (\ref{fittingformula}). As is described in Sec.\ III we first
fix a value of the exponent $\chi$ and then fit for the
other parameters. Repeating the fits in different sections of the
chain we seek the value of $\chi$ which makes the other
parameters the most stable as a function of $n$.
For $\Gamma=1.1$ the value of the exponent which yields the most
stable fitting parameters is $\chi \approx 0$.
From positions $n=50$ to $n=100$, and for both $h$, the fitted
correlation length $\xi$ agrees with the exact $\xi^{\rm ITF}$ with
a precision of 7 digits! It seems that, asymptotically,
the decay of $m^{\rm ITF}_{n}$ in the disordered phase is
purely exponential. This should be compared with the exponent
$\chi_{\rm bulk}=1/2$ characterizing the decay of the bulk
two-point correlation function in this phase.\cite{Bar-McC}
(For a similar difference in algebraic prefactors of one
and two-point correlation functions in
an $S=1$ chain see Ref.\ \onlinecite{Sor-Aff}.)
\begin{figure}[hbt]
\epsfxsize=\columnwidth\epsfbox{fig7b.eps}
\caption{ITF chain results. Ordered phase $\Gamma=0.9$. The local
values of the correlation length $\xi$ from Eq.\
(\protect\ref{fittingformula}), obtained
with different boundary fields $h$ and
exponents $\chi$, as a function of the position $n$ at $L=200$.
The exact correlation length $\xi^{ITF}$ is shown as a
horizontal line.
\label{fig:ITF2} }
\end{figure}
In the ordered phase for $\Gamma=0.9$ we can also start with
$\chi = 0$. In this case, however, the fitting parameters are not
stable,
and $\xi$ increases with $n$ for both values of $h$, as is
illustrated by Fig.\ \ref{fig:ITF2}.
This suggests that a nonzero prefactor is present. The value of
$\chi$ which makes the correlation length $\xi$ the most stable
seems to depend on the boundary field applied. We obtain
$\chi \approx 1.25 \pm 0.05$ for $h=10$ and $\chi \approx 0.9 \pm
0.05$ for $h=0.1$, we cannot see a universal prefactor. (Note
that for the bulk two-point functions the exponent of the
prefactor is $\chi_{\rm bulk}=2$ in this phase.) The values
of $\xi$ calculated with the above prefactors are shown
in Fig.\ \ref{fig:ITF2}. It is seen that using these prefactors we
find better estimates for $\xi^{\rm ITF}$ than with $\chi=0$.
Anticipating computational limitations in the investigation
of the $S=2$ cubic model in Eq.\ (\ref{model}),
we also analized shorter chains, such that
$L \approx 10 \xi^{\rm ITF}$ during the test calculations.
For $\Gamma=1.1$ and $L=100$
four digits of the exact $\xi^{\rm ITF}$ are recovered;
for $\Gamma=0.9$ and $L=50$ only rough estimates for
$\xi^{\rm ITF}$ are obtained, the error is around 10 \%.
Our observations can be summarized as follows:
\begin{itemize}
\item
The order parameter $m^{\rm ITF}$ can be calculated accurately
in both phases;
it can be decided whether the system is ordered or not.
\item
In the ordered phase, upper and lower bounds for
$m^{\rm ITF}$ are obtained by considering strong
and weak boundary fields, respectively.
\item
The correlation length $\xi^{\rm ITF}$
can be calculated very precisely in the disordered phase.
\item
In the ordered phase,
rather rough estimates of $\xi^{\rm ITF}$ can be found.
\end{itemize}
\end{appendix}
|
\section{Introduction}
Characterizing the impact of star formation on a galaxy's ISM is a
fundamental step towards understanding the interplay between star
formation and gas and, ultimately, the mechanisms responsible for
galactic star formation. In a large-scale star formation event,
stellar winds and supernova explosions from massive stars generate a
feedback mechanism by injecting energy into the ISM, which may produce
gas outflows and, in more extreme cases, superwinds (Elmegreen 1992,
Silk 1997, Shull 1993, Chu \& Kennicutt 1994, Heckman, Armus \& Miley
1990). Outflows/superwinds may act as a self-regulating mechanism for
the burst, by removing gas from the site of star formation
(Heckman 1997, Kennicutt 1989, Meurer et al. 1997). At a less extreme
level, OB associations will drive ionization and shock fronts through
the ISM, far away from the site of the massive stars, and in some cases
causing the star formation process to propagate spatially (Elmegreen
\& Lada 1977, McCray \& Kafatos 1987, Parker et al. 1992, Satyapal et
al. 1997, Puxley, Doyon \& Ward 1997). Exploring the interactions
between a starburst and its surroundings requires studies of
components of the host galaxy which are most directly affected by the
presence of the starburst. In the optical, the diffuse ionized gas
(DIG) is a prime candidate, and in this paper we investigate the
ionized ISM in two nearby starburst galaxies, NGC5253 and NGC5236.
The presence of DIG as a general component of the ISM of galaxies is
well known (e.g., Monnet, G. 1971). Over the last 15 years, a number
of studies have revealed its presence in a large variety of galaxy types:
spirals, star-forming irregulars and starbursts (Kennicutt \& Hodge
1986, Kennicutt, Edgar, \& Hodge 1989, Gallagher \& Hunter 1990,
Hunter \& Gallagher 1992, Hunter \& Gallagher 1997, Lehnert \& Heckman
1995, Rand 1998, Wang, Heckman, \& Lehnert 1998, 1999). Although
generally associated with the presence (or past presence) of massive
stars, the DIG can extend over 10 times larger spatial scales than the
ionizing stars (Reynolds 1991), and, as a result, the mechanism of
ionization of the DIG is still a matter of debate. Photoionization
from OB stars appears to account for the Reynold's Layer (Dove \&
Shull 1994) and may be a general mechanism for exciting the DIG
(Ferguson et al. 1996a, Ferguson, Wyse \& Gallagher 1996b, Hunter \&
Gallagher 1997). This requires that more than 20--30\% of the ionizing
photons leak out of HII regions. However, the emission line ratios of
the DIG are often very different from what is expected from
photoionization, and evidence has been accumulating in favor of mixed
photoionization/shock, or some other, heating mechanism for the DIG
(Sivan, Stasinska, \& Lequeux 1986, Hunter \& Gallagher 1990, Martin
1997, Rand 1998).
The DIG may be composed of at least two phases: a quiescent one, which
comprises 80\% of the H$\alpha$ emission and has a scale height
comparable to that of stars, and a turbulent one, with a scale height
about 3 times that of stars (Wang 1998, Wang, Heckman \& Lehnert
1997). These studies demonstrate that the effects of the massive
stars extend to galactic scales, and could affect the subsequent
evolution of the galaxy. Star formation in starburst galaxies
proceeds at a pace that is one-to-two orders of magnitude higher than
in ``quiescent'' galaxies and thus has a major impact on the DIG, as
seen in its kinematics, spatial distribution, and ionization (e.g.,
Marlowe et al. 1995). However, the details are not yet clear as to how
the properties of the DIG correlate with the star formation rate of
the host galaxy. For instance, does the importance of the DIG increase
in a starburst relative to a quiescent galaxy or does it simply become
brighter and, therefore, easier to observe (Wang et al. 1998)?
To date, studies of the physical conditions of the ionized ISM in
galaxies, and in starbursts in particular, have been pursued mainly via
long-slit spectroscopy of a limited number of regions. However,
structurally complex objects like starbursts cannot be unraveled
without fully accounting for the morphology of both gas and stars. A
complementary approach is to use narrow and broad band imaging to
obtain a complete map of the gas emission. While the advantage of
imaging is to fully characterize the spatial distribution of the
ionized gas, the disavantage is that only the brightest ionized lines
can be reasonably imaged with sufficient depth. In addition, the
matching of the narrow filter passband with the redshift of the galaxy
represents a technical difficulty for large samples of galaxies.
Imaging and long-slit spectroscopy provide
complementary approaches to the study of the ionized gas in galaxies.
Here we present images of the two nearby starburst galaxies NGC5253
(v=404~km/s) and NGC5236 (v=516~km/s) in the light of [OIII],
H$\alpha$, and [SII]. Variations of the intensity of ionization
lines, expecially the low ionization ones like [SII], are direct
indicators of changing physical conditions of the gas. The two
galaxies form a binary pair in the Centaurus group at a distance of a
few Mpc (4~Mpc for NGC5253, Sandage et al. 1994). They have completely
different characteristics: NGC5253 is a peculiar dwarf (Caldwell \&
Phillips 1989) and NGC5236 is a massive, grand design spiral,
classified as an SABc (Telesco et al. 1993); the two form a
metal~poor/metal~rich pair, with NGC5253 at about 1/6~Z$_{\odot}$ and
NGC5236 at about 2~Z$_{\odot}$. They are both experiencing a
high-intensity burst of star formation in their central regions,
possibly triggered by an encounter between the two about 1~Gyr
ago. This possibility was first suggested by van den Bergh (1980) on
the basis of various evidence, including the warping of the HI disk of
NGC5236 (Rogstad, Lockhart \& Wright 1974). Thanks to their closeness,
these galaxies are excellent laboratories for studying spatial
variations of the gas conditions in starbursts; 1$^{\prime\prime}$
corresponds to a linear scale of 19~pc in NGC5253, the size of a
typical HII region.
NGC5253 is a ``benchmark starburst'', with centrally concentrated
recent star formation superimposed on an older, quiescent stellar
population. The central star-forming region is very blue, although it
is crossed by dust lanes which produce patchy and heavy obscuration
and make this galaxy at the same time an excellent UV- and far-IR
emitter (Kinney et al. 1993, Aitken et al. 1982, Telesco et al. 1993,
Walsh \& Roy 1989, Calzetti et al. 1997). Radio observations reveal
that a large fraction of the most recent star formation is hidden by
dust (Turner, Ho \& Beck 1998). The bulk of the ongoing starburst is
located in an area 50--60~pc in size, where the stars are about 5~Myr
old. The UV emission in this area is dominated by a $\sim$3--4~Myr old
stellar cluster, but the ionization is being driven by a $\sim$2~Myr
old, dust-buried, central cluster (Calzetti et al. 1997, Crowther et
al. 1998). The star formation rate (SFR) density of
$\sim$10$^{-4}$~M$_{\odot}$~yr$^{-1}$~pc$^{-2}$ corresponds to the
maximum levels observed in star-forming galaxies (Meurer et
al. 1997). Extending beyond the starburst is a $\sim$300~pc region
where star formation has been active at a roughly constant level for
the last $\approx$100~Myr, with a SFR density about 0.01 of the
starburst's; a handful of bright stellar clusters with ages between 10
and 60~Myr are contained in this area (Calzetti et al. 1997). The
ionized gas around the starburst is slowly expanding, with a velocity
$\sim$10~km/s, at least within the inner 200~pc region (Martin \&
Kennicutt 1995, see also Strickland \& Stevens 1999). The HII emission
extends for $>$1~kpc from the center, with two identified kpc-scale
superbubbles in the western periphery, one of them expanding with a
velocity of 35~km/s (Marlowe et al. 1995). Moderately high
[OIII]/H$\beta$ ratios in the presence of high [SII]/H$\alpha$ ratios
up to $\sim$40~arcsec (800~pc) from the center suggest that shocks or
some mechanism other than photoionization contribute to the gas
excitation (from long slit spectroscopy, see Martin 1997). The
peculiar HI kinematic indicates rotation about the major axis of the
galaxy (Kobulnicky \& Skillman 1995), but an alternative
interpretation, that NGC5253 has accreted/is accreting relatively
unprocessed gas along the minor axis, has been suggested to account
for the unusually low CO luminosity (Turner, Beck \& Hurt 1997).
The starburst in NGC5236 is comparable in intensity to the spectacular
event in NGC5253. The central star formation extends for
$\sim$20$^{\prime\prime}$ across ($\sim$360~pc at 3.7~Mpc distance),
and is bright at all wavelengths, ranging from X-ray (Trinchieri,
Fabbiano \& Paulumbo 1985, Ehle et al. 1998), through the UV (Bohlin
et al. 1983, Kinney et al. 1993), optical, near-IR (Gallais et
al. 1991, Rouan et al. 1996), and mid-IR (Telesco et al. 1993), to the
radio (Turner \& Ho 1994). The nucleus proper is luminous in the
near-IR owing to the large dust obscuration in the center of the
galaxy (Gallais et al. 1991, Rouan et al. 1996). Large amounts of dust
are present, and form multiple dark lanes which surround the center
and cross it in the N-S direction. Gas inflow along the bar collecting
at the inner Lindblad resonance may be fueling the central starburst
(Petitpas \& Wilson 1998). An optically visible arc of star formation,
possibly the main source of the starburst's UV emission, lies about
6$^{\prime\prime}$ S and SW of the nucleus (Heap et al. 1993, Bohlin
et al. 1983). In HST-WFPC1 U-band observations, the arc breaks down
into a series of very young star clusters, with ages of
$\sim$2--6~Myr. The arc-shaped structure could be part of a ring
surrounding the nucleus where the other sections are not currently
actively forming stars (Gallais et al. 1991). Perhaps star formation
has not been co-eval, but sequential, propagating through the
ring. Differences between the Br$\gamma$ EW and CO EW maps support
this picture (Puxley et al. 1997). Mid-IR (Telesco et al. 1993) and
radio (Turner \& Ho 1994) maps depict a different picture. These show
two main sources amid diffuse emission: the northern-most source
coincides with an optical clump about 11$^{\prime\prime}$ NW of the
nucleus and slightly W of the central dust lane, while the southern
one is obscured at shorter wavelengths by the dust lane in which it
appears to be embedded (Telesco et al. 1993). The detection of only
two ``point-like'' sources at long wavelength led various authors to
advocate their extreme youth.
Whatever processes triggered the central starbursts in the two
galaxies, they were acting on two very different environments: a
grand-design, massive spiral galaxy in NGC5236, and a dwarf in
NGC5253. One candidate for the triggering perturbation is the
encounter between the two galaxies about 1~Gyr ago (van den Bergh
1980); if this is the case, the timing of the trigger is the
same. This would reduce by one the number of free parameters in the
problem. The difference between the two environments would then be the
major variable, making this pair an important test-bed for starburst
studies. This paper therefore focuses on the role of the host galaxy
environment on the evolution of the starburst by investigating the
physical properties and variations of the large-scale structure of the
ionized medium associated with each of the two starbursts. Section~2
describes the observations and data reduction; section~3 presents the
analysis of the observations, with special emphasis on nebular line
emission; the discussion is contained in section~4 and the conclusions
in section~5.
\section{Observations and Data Reduction}
Broad and narrow band images of NGC5253 (Figure~1) and NGC5236
(Figure~2) were obtained at the 2.5-m telescope of the Las Campanas
Observatory with the Direct Camera and a 2k$\times$2k CCD during the
nights of April 28 -- May 1, 1997. Broad band images were obtained
with 3~in.$\times$3~in. filters in the Harris U, V, and R. Narrow band
images were obtained using 2~in.$\times$2~in. filters on loan from
CTIO, centered at the redshifted wavelengths of
[OIII]$\lambda\lambda$4959,5007~\AA,
H$\alpha+$[NII]$\lambda\lambda$6548,6584~\AA~ and
[SII]$\lambda\lambda$6717,6731~\AA~ (see Table~1). The platescale on
the CCD is 0$^{\prime\prime}$.26/pix, implying a total field of view
of 8$^{\prime}$.8 for the broad band images. The small size of the
narrow band filter introduced vignetting at the edges of the CCD, and
the final unvignetted field of view was about 5$^{\prime}$.2; the presence
of scattered light from the edges of the filters further reduced the
useful field of view of the narrow band images to about
4$^{\prime}$.7. The seeing varied during the course of the four nights
in the range 0$^{\prime\prime}$.9--1$^{\prime\prime}$.2.
Given the surface brightness variation of more than a factor 100 from
the center to the edges of the starburst regions (and more than 10,000
in line emission intensity), the exposure times ranged from 30~s to
600~s in V and R and from 30~s to 1,200~s in U and in the narrow band
filters, to achieve suitable exposure levels in different regions
of the galaxies. Offsets of a few arcseconds between frames were
introduced to remove cosmetic defects (bad pixels and two central bad
columns) from the final combined images. Table~2 lists for both
galaxies the total exposure time in each filter.
Data reduction followed the standard procedure of bias subtraction,
flat-fielding, registration, and co-addition of the images. Both dome
and twilight exposures were used to remove pixel-to-pixel variations
and illumination patterns from the images. Residual scattered light in
the [SII] filter was removed by subtracting a surface fit to the
background. The background fit for NGC5253 was adopted for both
galaxies, since the background of NGC5236 could not be fit as this
galaxy completely fills the field of view. The two central bad columns
of the chip were linearly interpolated in each image with values from
surrounding columns. Cosmic rays were removed from individual frames
before co-addition, using an algorithm developed by M. Dickinson
(1997, private communication) for the identification of sharp,
positive discontinuities over scales of $\sim$1~pix. This technique
removed around 80--90\%~ of the cosmic rays; final co-addition of
multiple frames removed most of the remaining events. Both galaxies
were observed close to their culminating points, and the effects of
airmass variations were generally less than 3\%; exception were U and
[OIII], where such effects were as large as 12\%~ and 5\%,
respectively, and corrections were therefore applied. Absolute
calibrations were obtained from observations of two spectrophotometric
standards from Hamuy et al. (1994). One of the standards was also
observed during the night at different azimuths to derive airmass
corrections. Absolute flux calibrations are listed in Table~1 for all
filters, together with the internal error (in percentage).
\subsection{Emission Line Images}
More than one emission line is included in each of the three narrow
band filters. Both the redshifted [OIII]$\lambda$4959~\AA~ and
[OIII]$\lambda$5007~\AA~ contribute to the emission in the 5000/70
filter; the second line is located almost at the center of the
passband, while the [OIII]$\lambda$4959~\AA~ is located on the ramp,
making the determination of its contribution rather uncertain. Our
estimates give a best value of $\sim$95\% for the filter transmission
at the redshifted [OIII]$\lambda$4959~\AA~ relative to the filter
transmission at [OIII]$\lambda$5007~\AA, with a range between 75\% and
106\%. Three lines, the redshifted H$\alpha$, [NII]$\lambda$6548~\AA~
and [NII]$\lambda$6584~\AA, contribute to the emission in the 6563/78
filter, with the reddest [NII] line located at 90\% and 72\% of the
peak transmission for NGC5253 and NGC5236, respectively. Both the
redshifted [SII]$\lambda$6717~\AA~ and [SII]$\lambda$6731~\AA~ are
located close to the transmission peak in the 6737/76 filter.
The calibrated V and R images were used to subtract the stellar
continuum from the [OIII], H$\alpha$+[NII], and [SII] images. After
matching the FWHM of the stars in the broad and narrow band frames,
the continuum images were recursively rescaled and subtracted from the
narrow band images till optimal removal of the galaxy stellar
continuum was achieved. The field stars were initially used to obtain
a first guess on the scaling factor, but refinements on this factor
were necessary due to the bluer stellar continuum of the galaxies'
centers relative to the stars. For the weak [SII] emission we first
subtracted the H$\alpha$+[NII] nebular emission from the R-band image
and then used this nebular~emission-free image to remove the stellar
continuum from the [SII] image. There is a marked color gradient in
the V-band image of NGC5236 with the center bluer than the external
regions; therefore accurate continuum subtraction over the entire
[OIII] image of NGC5236 could not be achieved. We paid particular
attention to the central region, and obtained a satisfactory [OIII]
emission~line image of the inner $\sim$40$^{\prime\prime}$. This
region is comparable in size to the area of H$\alpha$ emission
detected above 5$\sigma$ (see section~3.3 below), and thus is sufficient
for our purposes.
The accuracy of the calibration of the narrow band filters was checked
against ground-based spectrophotometry of the centers of the galaxies
(Storchi-Bergmann, Kinney \& Challis 1995) and, for NGC5253, against
H$\alpha$ images obtained with the Hubble Space Telescope Wide Field
and Planetary Camera 2 (Calzetti et al. 1997). In all cases, our
calibrations gave flux values slightly higher than the spectra and the
HST image. For NGC5253, our [SII], H$\alpha+$[NII] and [OIII] fluxes
are about 5\%, 7\%~ and 10\%~ larger, respectively, than what is
measured from the spectrum. For NGC5236, the [SII] and H$\alpha+$[NII]
image fluxes are 8\% and 2\% larger, respectively, than the
spectrum. To compare our H$\alpha+$[NII] image of NGC5253 with the HST
one, the contribution of the [NII]$\lambda$6584~\AA~ line had to be
removed; from the spectrum we estimate that the [NII] flux is on
average 13\%~ of the H$\alpha$ flux, although variations are expected
as a function of position (Kobulnicky et al. 1997). After the
subtraction of this contribution, our narrow band image gave fluxes
consistently $\sim$6\% higher than the HST image. The source of this
fairly small, but systematic discrepancy is unclear. We have
considered under-subtraction of the stellar continuum, filter
calibration, and presence of Balmer absorption in the calibration
stars, but none of those reproduces all of the observed
discrepancy. However, the relative calibration of the three narrow
band images is good at the 5\%~ level. For consistency, we rescale our
emission line images to the spectroscopic/HST values.
For NGC5253, HST WFPC2 images centered on the H$\beta$ line emission
(Calzetti et al. 1997) are used here to supplement the ground based
images. The HST H$\beta$ image has been rotated, smoothed, resampled,
and registered to match the ground-based images.
\section{Analysis and Results}
\subsection{Line Ratio Maps}
Ratios of metal-to-hydrogen lines are common diagnostics of the
physical conditions of the ionized gas. We produced maps of
[SII]$\lambda\lambda$6717,6731\AA/H$\alpha$ and
[OIII]$\lambda$5007\AA/H$\beta$ for NGC5253 (Figures~3c, 3e and 3f),
and [SII]$\lambda\lambda$6717,6731\AA/H$\alpha$ and
[OIII]$\lambda$5007\AA/H$\alpha$ for NGC5236 (Figure~4e and 4f). The
line ratio maps have been created using a 5~$\sigma$ detection
threshold for each line, after all the images have been smoothed to the
seeing of the photometrically worst night ($\sim$1$^{\prime\prime}$.2)
and resampled to 5~pix$\times$5~pix bins
(1$^{\prime\prime}$.3$\times$1$^{\prime\prime}$.3). The [OIII] line
maps have been divided by a factor 1.3 to remove the contribution from
the [OIII]$\lambda$4959\AA. For NGC5236, we don't have an H$\beta$
emission line image, thus the corresponding ratio map
[OIII]$\lambda$5007/H$\beta$ could not be constructed. We use the
H$\alpha$ image instead, with cautionary remarks about the potentially
large effects of reddening variations in the center of the galaxy (see
below). We also note that the difficulty of subtracting the continuum
from the [OIII] image for this galaxy contributes to the larger
uncertainty in the line fluxes; a number of the data bins are below our
required 5~$\sigma$ threshold, and the `usable'
[OIII]$\lambda$5007/H$\alpha$ map includes the central
$\sim$30$^{\prime\prime}$ region only. This is slightly, but possibly
significantly, smaller than the extent of the 5~$\sigma$ H$\alpha$
ionized region, which occupies the central $\sim$40$^{\prime\prime}$
(see section~3.3). In NGC5253, regions beyond the central
$\sim$30$^{\prime\prime}$ in radius have H$\beta$ flux detections
below 5~$\sigma$, and we have used the [OIII]/H$\alpha$ ratio instead,
assuming that reddening corrections are small at large distance from
the center of the starburst (see discussion below).
\subsubsection{Underlying Stellar Absorption}
Corrections for the stellar absorption underlying the Balmer lines are
important expecially at the faintest surface brightness levels, where
the EW of the emission line is correspondingly small. We used the
ratio of the H$\alpha$ and the H$\beta$ emission to the corresponding
underlying continua to derive maps of the EW of these lines; for
H$\alpha$ we used the R-band image as continuum, while for H$\beta$ we
used the extrapolated continuum image from the HST V and I images of
NGC5253 (Calzetti et al. 1997). The line fluxes were then corrected
for the presence of underlying stellar absorption with constant value
EW=3~\AA~ (e.g., McCall et al. 1985). Figures~3d and 4d show the
H$\alpha$ EW emission maps of the two galaxies, after the
correction. Unknown variations of the underlying stellar
absorption EW increase the undertainty in the line flux at the detection
threshold, where the emission line EWs are generally small (Figures~3d
and 4d). In addition, the contribution of an intermediate/old
underlying stellar population proportionally increases as the distance
from the center of the starburst (the young population) increases,
thus gradually changing the underlying stellar absorption from
$\sim$3~\AA~ to $\approx$5~\AA. The combination of the two effects
implies an uncertainty of $\sim$20\% and $\sim$50\% for the H$\alpha$
measurements at the detection threshold in NGC5253 and NGC5236,
respectively. The variable underlying stellar absorption is taken into
account in the following sections every time its effect is relevant to
the measurements.
\subsubsection{Dust Reddening}
Corrections for dust reddening are generally small for the
[SII]/H$\alpha$ and [OIII]/H$\beta$ maps, due to the closeness in
wavelength of each pair of lines, but can be large for the
[OIII]/H$\alpha$, because the wavelength difference is large. We
discuss reddening corrections separately for each of the two galaxies.
Dust extinction in the central
$\sim$30$^{\prime\prime}\times$30$^{\prime\prime}$ of NGC5253 is
highly variable (Calzetti et al. 1997); there is a E-W dust lane
bisecting the central section of the galaxy and the central ionizing
stellar cluster is deeply embedded in a highly opaque dust cloud. We
thus use the HST reddening map of Calzetti et al. (1997) to remove
extinction effects from the line ratios. We assume the reddening is
foreground, which should be a reasonable approximation for most
regions, since the H$\alpha$/H$\beta$ ratio approaches the unreddened
case in the vast majority of the bins. However, we already know that
the foreground geometry is altogether wrong in the case of the central
cluster, where the amount of extinction is above A$_V$=10~mag and the
geometry is known not to be foreground (e.g., Beck et al. 1996); such
cases should be statistically insignificant when trends between lines
are analyzed, as they include a relatively small number of
bins. Regions beyond $\sim$30$^{\prime\prime}$ radius, where the HST
reddening map is not available, are corrected with the assumption of a
small, constant reddening E(B$-$V)=0.1, of which 0.05 are from our
Galaxy (Burstein \& Heiles 1982). This assumption is reasonable, as
the H$\alpha$/H$\beta$ map indicates that the reddening decreases to
small values beyond a radius of $\sim$20$^{\prime\prime}$ from the
center (Calzetti et al. 1997).
In the absence of a reddening map for NGC5236, we have adopted the
constant value E(B$-$V)=0.35 for the dust extinction correction, which
includes both intrinsic and Galactic foreground extinction, as derived
from ground-based spectroscopy of the central starburst (Calzetti et
al. 1994). Given the small wavelength difference between [SII] and
H$\alpha$ the impact of reddening corrections is no larger than 6\%
for a reddening variation between E(B$-$V)=0 to 0.7. The impact is of
course much larger for the [OIII]$\lambda$5007\AA/H$\alpha$; the
intrinsic ratio changes by 40\% if E(B$-$V)=0.7 instead of 0.35. We
know from the study of Telesco et al. (1993) that there is a
consirable amount of dust with a complex geometry in the center of
NGC5236. For this reason, the [OIII]/H$\alpha$ ratio map, which will
be briefly discussed in the next sections, should be considered a
preliminary substitute for [OIII]/H$\beta$ for this galaxy.
\subsubsection{The Contribution of [NII] to the H$\alpha$ maps}
The ratio of the HST/ground~based H$\alpha+$[NII] images gives an
estimate of the [NII] intensity change across the central region of
NGC5253, because the HST image does not contain the
[NII]$\lambda$6584~\AA~ emission while the ground~based image
does. The image ratio appears constant to within 12\% in the central
$\sim$27$^{\prime\prime}$ (radius), which is where the S/N is
high. Thus variations of the [NII] intensity are not expected to be
more than twice its average value. This is in agreement with results
from long-slit spectroscopy (Lehnert \& Heckman 1995, Martin 1997),
which indicate that the variation in the [NII]/H$\alpha$ ratio is very
small for the central $\sim$25--30$^{\prime\prime}$ of NGC5253 (see,
however, Kobulnicky et al. 1997). In the light of this result, we
derived a ``pure'' H$\alpha$ image by removing 17\%~ of the flux from
the original image. The 17\%~ figure represents the contribution of
the two [NII]$\lambda\lambda$6548,6584~\AA~ lines.
For NGC5236, the contribution of [NII]$\lambda\lambda$6548,6584\AA~ to
the H$\alpha$ map has been estimated from spectroscopic data
(Storchi-Bergmann et al. 1995): the [NII] line contributes $\sim$39\%
of the total line flux in the 6563/78 filter, in agreement with the
high metallicity of the galaxy.
In the shocked regions of both galaxies, our derived [SII]/H$\alpha$
is a lower limit to the true value as [NII]/H$\alpha$ is expected to
increase for increasing [SII]/H$\alpha$, leading to an underestimate
of the [NII] contribution to our H$\alpha$ maps. This will have in
general the effect of weakening the shock diagnostics, an opposite
effect to that induced by variable underlying stellar absorption.
\subsection{NGC 5253}
\subsubsection{The Ionized Gas Morphology}
The morphology of the nebular gas emission in NGC5253 has been
described by a number of authors (e.g., Marlowe et al. 1995, Martin \&
Kennicutt 1995, Calzetti et al. 1997). We review here a few basic
facts. The ionized gas emission is circularly symmetric around a
stellar cluster located almost at the geometric center of the galaxy
(Figure~1, bottom panel); this cluster, with an age of $\sim$2~Myr, is
also the youngest stellar cluster in the galaxy (Calzetti et al.
1997). The azimuthally-averaged H$\alpha$ emission monothonically
decreases in surface brightness from the cluster outward. The regular
morphology of the gas emission is in striking
contrast with the morphology of the UV and optical stellar continuum
(Figures~3a and 3b), which is elongated from NE to SW, along the major
axis of the galaxy (e.g. Martin \& Kennicutt 1995). We clearly detect
in each of the H$\alpha$, [OIII] and [SII] maps the two western bubbles
described in Marlowe et al. (1995): the one closer to the minor axis
is the weakest of the two, and we detect the outer shell of the
expanding gas; the other, which is almost along the major axis of the
galaxy, is well detected and shows a wealth of substructure (Figures~1
and 3). A number of filaments extend outward from the center, both in the
North and in the East-South direction. The presence of ionized
gas along the dust lane, SE of the center, is detected in both our
H$\alpha$ and [OIII] images, with a hint in the 5~$\sigma$ [SII] image.
\subsubsection{Photoionization and Shock-Ionization}
Figure~5 (panel~a) shows the line ratios measured in each
1$^{\prime\prime}$.3 resolution element and compares those with models
for gas photoionization and for shock excitation. The photoionization
models give the variation of the line ratios for changing ionization
parameter U. One set of models has been taken from Martin (1997), who
ran CLOUDY (Ferland 1993) for a range of metallicities and effective
temperatures of the ionizing source. Two other models are from
Sokolowski (1993), who analyzed the cases of depleted metal abundances
and of hardened photoionizing continuum; the models assume cosmic
abundances and an ionizing source given by an instantaneous burst of
star formation with a Salpeter stellar mass function up to
120~M$_{\odot}$. The last model reproduces the scenario in which the
soft ionizing photons are the first to be absorbed by the ISM, thus
the ionizing continuum hardens as it travels across the
galaxy. Predicted line ratios for shock excitation are from Shull \&
McKee (1979), for cosmic abundances and a range of shock velocities
and for the special case of depleted metal abundances.
The metal-to-hydrogen line ratios change as a function of the
ionization parameter and this can potentially explain the observed
variation in Figure~5 (e.g, Hunter 1994). The ionization parameter U
measures the relative amount of ionizing photons relative to the
amount of gas. Increasing the distance from the ionizing source
decreases the value of U, lowering the [OIII]/H$\beta$ ratio and
increasing the [SII]/H$\alpha$ ratio (Domg\"orgen \& Mathis 1994). The
data of NGC5253 appear to follow this trend qualitatively both in
Figure~5 and in the ionization map [OIII]/[SII] of Figure~3g. Except
along the dust lane (see discussion below), the [OIII]/[SII] ratio
decreases from the center to the edges of the ionized region. However,
a quantitative comparison (Figure~5, panel~a) shows that the
[OIII]/H$\beta$ value decreases less steeply than expected from
variations of the ionization parameter, a trend already noted by
Martin (1997) for a sample of dwarf galaxies.
The photoionization model with T$_e$=50,000~K and
[O/H]=0.2~[O/H]$_{\odot}$, which closely matches the metallicity of
NGC5253 ($\sim$1/6~[O/H]$_{\odot}$), marks a lower envelope to the
data points in Figure~5, while it is in reasonable agreement with the data
at the highest values of U, in regions closest to the central
ionizing source in the starburst. A somewhat better representation of
the data is given by the model with depleted abundances (Sokolowski
1993), but most of the data points are still above the locus of the
photoionization lines. The observed line ratios behave as if there is
an increasingly important shock component (Martin 1997) or the
radiation spectrum is progressively hardened towards the external
regions (Wang 1998).
Ionization in NGC5253 can be directly compared with the well-studied
Large Magellanic Cloud. Figure~5, panel (b), shows the [OIII]/H$\beta$
versus [SII]/H$\alpha$ line ratios of a sample of HII regions, giant
and supergiant shells in the LMC from Hunter (1994). Although some of
the data points show the same extreme behavior as NGC5253, the
majority of the shells agree with photoionization models. The
metallicity of the LMC is about twice that of NGC5253, thus the
comparison between the two galaxies is not immediate, as the LMC data
naturally occupy a locus to the lower left relative to the NGC5253
data. Neverthless, the majority of the LMC line ratios are between the
solar metallicity and the depleted abundances models, and the
strongest outliers are in giant shells, which seem to require a
hardened radiation field.
To further discriminate between ionization mechanisms, we have plotted
the ratios [SII]/H$\alpha$ and [OIII]/H$\beta$ as a function of the
physical distance from the central star cluster in NGC5253 (Figure~6).
The mean value of [SII]/H$\alpha$ ([OIII]/H$\beta$) increases
(decreases) for increasing distance from the `center of
ionization'. Again, photoionization models reproduce qualitatively,
but not quantitavely, this trend. The relationship between ionization
parameter and distance has been taken from Martin (1997, see her
Equation~1); the size of the ionized sphere has been assumed to
correspond to the size of the H$\alpha$ emission, around
71$^{\prime\prime}$--81$^{\prime\prime}$ in radius, or 1.4--1.6~kpc. A
quantitative test shows that photoionization alone cannot fully
explain the observed trend of the line ratios, and, in particular,
cannot account for the increasing {\em spread} about the mean values
with increasing distance. The increasing spread with distance is quite
evident in the [OIII]/H$\beta$ diagram (Figure~6b). We interpret this
as an effect of the increasing importance of shock-ionization (or
other non-photoionization mechanism, see Haffner, Reynolds \& Tufte
1999) over photoionization further from the center. The physical
extent of the starburst and metallicity variations may play a role in
the line ratio spread, but we do not expect these to be the dominant
effects. We will show in the next section that the starburst
population extends over a much smaller area, less than 1/6, than the
ionized gas. Both the mean value and the spread of [SII]/H$\alpha$
increase for decreasing H$\alpha$ surface brightnesses (Figure~7, see
Wang, Heckman \& Lehnert 1998, Martin 1997, and Ferguson et al. 1996b
on a variety of galaxies), supporting what is observed in Figure~6.
The reddening-corrected ionizing photon rate from the starburst is
$\log$~Q(H$^o$)=52.57--52.78, depending on the dust opacity adopted
for the central star cluster (9~mag$\le$A$_V\le$35~mag, Calzetti et
al. 1997). These values correspond to a Str\"omgrem radius
R$_S$=240--280~pc, for an electron density of 93~cm$^{-3}$ and
temperature $\sim$10,000~K (Storchi-Bergman, Kinney \& Challis 1995,
CKS94), and for a filling factor of 0.01 (Martin 1997). The calculated
Str\"omgren radius is at least a factor $\sim$4.5 smaller than the
extent of the H$\alpha$ emission.
\subsubsection{The Morphology of `Shocks' and DIG}
Adopting arbitrarily the constraints $\log([OIII]/H\beta)>0.2$ and
$\log([SII]/H\alpha)>-0.35$ to discriminate between predominance of
photoionization and predominance of shock excitation/hardening of
radiation (or other mechanism), the location of the `shocked' regions
is graphically represented in Figure~3h. The figure shows that the
purely photoionized region has a circularly symmetric shape centered
almost exactly on the main ionizing cluster, with radius
$\simeq$560~pc; it is comparable in size to the stellar population of
the starburst, although the morphology of the two is different (see
next section). The `shocked' gas has a markedly asymmetric morphology;
the majority of the bins with $\log([OIII]/H\beta)>0.2$ and
$\log([SII]/H\alpha)>-0.35$ is located in the south-west region, and
there appears to be an overlap of filaments and arches extending out
of the main starburst area. The `shocked' gas extends in the direction
of the major axis of the galaxy; one would expect expanding gas to
prefer the direction of the minor axis (e.g., Heckman et al. 1990, De
Young \& Heckman 1994, Martin 1998, Meurer, Staveley-Smith, \& Killeen
1998), which does not seem to be true for NGC5253.
The constraint $\log([SII]/H\alpha)>-0.35$ corresponds to a H$\alpha$
surface brightness less than
6.39$\times$10$^{-16}$~erg~cm$^{-2}$~s$^{-1}$~arcsec$^{-2}$, or a
normalized surface brightness SB(H$\alpha$)/SB$_e<$0.01, in
Figure~7. We stress again here that variations in the underlying
stellar absorption would have typically no more than a 20\% effect on
both the SB(H$\alpha$) and the [SII]/H$\alpha$ ratio (see
Figure~3d). The ratio SB(H$\alpha$)/SB$_e<$0.01 marks a sharp increase
in the median value of [SII]/H$\alpha$. A similar sharp break is
evident also in the histogram of the bins with specific value of
H$\alpha$ surface brightness (Figure~8): below
SB(H$\alpha$)/SB$_e=$0.01 there is a large gradient in the relative
number of bins. Such ``breaks'' have been pointed out by Wang (1998)
as the transition between HII regions and DIG. In the case of NGC5253,
the DIG surrounds the central starburst up to a distance of at least
$\sim$1.4--1.6~kpc in some directions (to our detection limit).
\subsubsection{The Morphology of the Starburst Population}
In order to compare in detail the morphologies of the ionized gas and
of the ionizing stars, we have removed the underlying galaxy from the
broad band image, so that the structure of the current starburst would be
enhanced.
The U band image includes the [OII]$\lambda$3727~\AA~ doublet emission
in its passband. In the central ~15$^{\prime\prime}$ [OII] has
EW=130~\AA~, thus providing about 20\% contribution to the U
emission. Since we did not observe the [OII] emission, we used the
[SII] map, which was rescaled to the [OII] intensity observed by
Storchi-Bergmann et al. (1995). This procedure relies on the
(reasonable) assumption that [OII] and [SII] have the same morphology
and intensity distribution; this is supported by the observations of
Martin (1997). The H$\alpha$ contribution to the R band image has been
removed in a more straightforward manner (see section~2). All three
broad band images were corrected for the effects of dust reddening
using the HST reddening maps (or the constant value E(B$-$V)=0.1
outside the range of the HST maps) and the prescription of Calzetti et
al. (1997).
The U-band isophotes external to $\sim$70$^{\prime\prime}$ follow an
exponential profile (Caldwell \& Phillips 1989), typical of the old
stellar populations in spheroidal and irregular dwarf galaxies.
Indeed, NGC5253 would very likely be classified as a dwarf elliptical
(Sersic et al. 1972) if it weren't for the central starburst. At
smaller radii there is a clear excess relative to the exponential fit;
Caldwell \& Phillips attribute this excess to the star formation event
which occurred in the galaxy over the last $\approx$1~Gyr. Not all of
this excess is composed of ionizing stars; along the galaxy major
axis, the region between 50$^{\prime\prime}$ and 70$^{\prime\prime}$
is not associated with H$\alpha$ emission (to our detection
limit). The ionizing starburst appears more concentrated than the
population excess over the exponential light profile. To remove the
non-ionizing stellar population underlying the ionizing starburst, we
used the isophotes between 50$^{\prime\prime}$ and
70$^{\prime\prime}$, and attempted both an exponential profile and a
r$^{1/4}$ law model. The latter fits the non-ionizing population
isophotes to r$\sim$34$^{\prime\prime}$ better than the exponential
profile. This isophotal profile was extrapolated to the center and
subtracted from the original image. The residual, namely the central
starburst, is shown in Figure~3a for the U band. All three continuum
images present the same morphology, thus hinting that dust extinction
does not affect the global appearance in the optical passbands. The
colors of the underlying galaxy are fairly uniform, with values
U$-$V=0.5$\pm$0.2 and V$-$R=0.80$\pm$0.25, typical of a stellar
population dominated by A7 and later type stars, which will not
contribute to the photoionizing luminosity.
The comparison between the starburst continuum emission and any of the
line emission or line ratio maps shows an obvious characteristic: the
line emission is more extended, by more than a factor $\sim$2, than
the continuum emission (see Figure~3a with 3b). This is true for both
the photoionized and shock-ionized parts of the nebular line emission,
confirming that the photoionized gas is displaced relative to the
ionizing stars by more than $\sim$1~kpc from the external perimeter of
the starburst. A plot of the H$\alpha$ surface brightness as a
function of the U$-$V color of the starburst population shows the
expected trend that higher SB(H$\alpha$) coincide on average with
bluer U$-$V colors (Figure~9). The regions with
SB(H$\alpha$)/SB$_e>$0.01 have typical colors
U$-$V$\simeq-$0.5,$-$1.3, corresponding to ages between 1 and 100~Myr
for constant star formation and between 1 and 30~Myr for an
instantaneous burst population (Leitherer \& Heckman 1995); this
agrees with the age range found by Calzetti et al. (1997). A few
points in this region have U$-$V$<-$1.6, bluer than the bluest models
for stellar populations. This reflects uncertainties in the color
derivation and, possibly, an imperfect subtraction of the strong [OII]
emission from the U-band image. Lower SB(H$\alpha$) correspond to
regions with typical colors of nonionizing or mildly ionizing
populations. We highlight again that the ionizing stellar population
extends over an area which is $<$1/6 of the area of the detected gas
emission.
\subsubsection{Star Formation in the Dust Lane}
The values of [OIII]/H$\alpha$ along the dust lane (see the little
`horn' sticking out at the bottom left of Figure~3f) have median
$\simeq$6, compatible with the values in the center of the
starburst. In addition, the ratio [OIII]/[SII] remains high, around or
above 10 (about 1/2 the value of the central cluster, see Figure~3g),
along the entire dust lane. Insufficient reddening correction due to
the presence of the dust lane would make both ratios even higher. This
is one of the areas responsible for the marked spread in the
[OIII]/H$\beta$ values at large distance from the center. We can place
an upper limit [SII]/H$\alpha<$0.35 in this area. Both line ratios are
compatible with this region being almost purely photoionized. However,
there are no obvious ionizing stars in this area, although we cannot
exclude that star formation is heavily embedded in the dust lane, and
thus is not visible. Even if this is the case, star formation in the
lane is happening at a relatively low intensity level; the
star-formation-sensitive 10~$\mu$m map of Telesco et al. (1993),
indeed, does not show emission along the dust lane, and dust
obscuration is less effective at 10~$\mu$m than in the optical.
\subsection{NGC 5236}
\subsubsection{Morphology of the Starburst}
The global morphology of the ionized gas emission in NGC5236 is far
simpler than in NGC5253, and, likely, easier to interpret. Most of the
H$\alpha+$[NII] emission comes from the central
$\sim$40$^{\prime\prime}$, where the starburst is located, and along
the spiral arms (Figure~2b). Unlike NGC5253, there is little or no
evidence for arcs, loops or filaments of ionized gas extending outward
from the central starburst. In the center of the galaxy, the brightest
part of the H$\alpha+$[NII] emission, above 15$\sigma$, occupies a
region $\sim$30$^{\prime\prime}$ across (corresponding to a physical
size of 540~pc), comparable in size and morphology to the bright blue
stellar emission detected in the U band (above 50$\sigma$, Figures~4a
and 4b). The optically brightest part of the starburst is located in
the south-western arc of blue stellar clumps, about
15$^{\prime\prime}$ in length. The northern tip of the arc appears to
bend in the east direction, but this morphology probably is an effect
of the crossing of the dust lane (see Gallais et al. 1991). The
arc-shape of the stellar continuum is fairly well mirrored by the
H$\alpha+$[NII] emission, with no obvious exceptions. The ionized gas
and blue star morphology of the center of NGC5236 is typical of the
central starbursts hosted in massive galaxies, where rings, arcs, and
``spirals'' of star formation are common structures (Maoz et al. 1996,
Colina et al. 1997). All characteristics of NGC5236, including those
described below, are consistent with star formation occuring in a
sharply bounded inner nuclear disk, perhaps defined by the inner
Lindblad resonance (ILR) as suggested by Telesco et al. (1993).
Figure~4c displays the HST-WFPC1 image of the center of the galaxy
in the F336W filter (Heap et al. 1993, roughly corresponding to the
U-band), where the arc (`A' in Figure~4c) clearly splits into several
individual stellar clusters and the northern `bend' of the arc (`B' in
Figure~4c) splits into three clusters. We cannot resolve the
individual line emission of each of the clusters in `B' from the
ground-based image, but their summed flux locates the peak of
H$\alpha+$[NII] emission in the galaxy center, with a total observed
flux F(H$\alpha+$[NII])=1.40$\times$10$^{-12}$~erg~s$^{-1}$~cm$^{-2}$,
measured in an aperture of 2.6$^{\prime\prime}$ diameter. The nucleus
(`N' in Figure~4c) is located about 6$^{\prime\prime}$ NE of the arc;
it appears as a lump in the ground-based U-band image (Figure~4a), but
with a weak H$\alpha+$[NII] emission (Figure~4b).
About 11$^{\prime\prime}$ North-West of the arc there are two bright
HII knots (`C' and `D' in Figure~4c; clump `D' is not visible in the
U-band image); they have comparable intensity in the narrow
line emission, with the northern-most of the two (`D') being only 30\%
brighter than the other, but very different U brightnesses, with `D'
being 5.9 times fainter than the other. We identify `D' as coincident
with the mid-IR northern source (Telesco et al. 1993). The difference
in line/continuum emission between the two knots is likely an age
effect with `D' being younger than `C'; if it were simply an effect of
dust reddening `D' would appear in near-IR imaging, but this is not
the case (Gallais et al. 1991). The younger age of knot `D' is
supported by the value of the H$\alpha$ EW, which is about 180~\AA~
for this knot, while it is only about 90~\AA~ for knot `C'. Larger
values of the EW(H$\alpha$) locate comparatively younger regions
(Leitherer \& Heckman 1995); in the case of the center of NGC5236, an
imaginary line joining region `B' with knots `C' and `D' identifies
the youngest part of the starburst, with values
EW(H$\alpha$)$\approx$200~\AA~ (Figure~4d), about twice those of
surrounding regions (Telesco et al. 1993). The inferred ages from such
EWs are less than 10$^7$~yr, for an instantaneous burst of star
formation (Leitherer \& Heckman 1995, see Puxley et al. 1997).
The three luminous condensations in the arc (`A' in Figure~4c,
corresponding to multiple clusters in the HST image) are between 1.61
and 2.05 times brighter in U than clump `B', while they are a factor
between 2.4 and 4.2 fainter in H$\alpha+$[NII]. In [OIII], the features
in arc `A' are about 2.4 times brighter than `B'. If the metallicity
along the arc is roughly constant, these differences are immediately
understandable in terms of dust reddening, with `B' being more
reddened than `A'. This is reasonable as `B' is located very
close to the NS dust lane.
\subsubsection{Gas Excitation}
The [OIII]$\lambda$5007\AA/H$\alpha$ ratio is plotted as a function of
[SII]$\lambda\lambda$6717,6731\AA/H$\alpha$ in Figure~10.
Sokolowsky's models, derived for cosmic abundances, are expected to
work fairly well for this galaxy, whose center has average metallicity
about twice solar. The data are not inconsistent with photoionization
models, in the entire range considered. There is little evidence for
shocks in NGC5236, although our line ratios cannot be used as the only
criterion for deciding the ionization mechanism, because of the
potential for heavy dust reddening effects in the [OIII]/H$\alpha$
ratio. The plot of [SII]/H$\alpha$ as a function of the distance from
the H$\alpha$ peak (Figure~11a) also shows that photoionization
appears to be the main gas excitation mechanism, as the trend of the
upper envelope to the points closely follows Sokolowski's model for
depleted abundances. Further support to the photoionization picture
comes from the range of values covered by the [SII]/H$\alpha$ ratio:
it is very close to that measured in NGC5253, despite the fact that
NGC5236 is at least one order of magnitude more metal-rich
(cf. Figure~11a with 6a). The latter conclusion does not qualitatively
change even if there is a 50\% uncertainty in the stellar absorption
underlying the H$\alpha$ emission or a similar uncertainty in the
[NII] contribution to the H$\alpha$ image.
The plot of the [OIII]/H$\alpha$ ratio as a function of the distance
from the peak of the H$\alpha$ emission is instead fairly inconclusive
(Figure~11b): here the scatter in the data points dominates any
trend. The scatter in Figure~11b is likely the superposition of two
effects: one is the inhomogeneity of the dust reddening, which we
cannot correct for with our data, the other may be the lack of a
correlation between the line ratio and the distance from the peak of
the H$\alpha$ emission. The presence of the second effect is confirmed
by Figure~11a. In this case variations in the reddening induce small
changes in the line ratio; neverthless, the plot still shows a fairly
large scatter. The most straightforward interpretation is that the
peak of the H$\alpha$ emission is not the absolute peak of the ionized gas
emission. Unlike the case of NGC5253 (Figure~6 and discussion in
Calzetti et al. 1997), the gas morphology in the center of NGC5236
cannot be described as the effect of a main central ionizing source,
but is far more complex with multiple emission peaks of almost
comparable intensity (see Figure~4b).
As in NGC5253, the largest values of the [SII]/H$\alpha$ ratio are
reached in the regions of lowest H$\alpha$ surface brightness
(Figure~12). Here, however, the scatter is much larger than in the
case of NGC5253, probably due to the insufficient extinction
correction of the H$\alpha$ surface brightness and uncertain
correction for the underlying stellar absorption. Also, the H$\alpha$
surface brightness limit reached for NGC5236 is about 3 times higher
than for NGC5253, due to shorter exposure times in both the 6563/78
and the R band filters, only partially compensated by the fact that
the red continuum of NGC5236 is about 5 times brighter than that of
the other galaxy.
The histogram of the number of area bins having a specific value of
the H$\alpha$ surface brightness (Figure~13) shows that the two
galaxies have similar behavior (slope and upper limit) at the high
brightness end, but differ quite substantially in the low surface
brightness regime. In particular, NGC5236 does not show the `break' in
the power-law trend shown by NGC5253. Thus, the transition between HII
regions and DIG is less clear in the spiral galaxy. Such variety of
behaviors among galaxies has been previously observed (Wang, Heckman \&
Lehnert 1998).
The starburst population of NGC5236 is not easily separated from the
underlying stellar population, because of the presence of uneven
dust/stellar population distribution across the entire galaxy. The
direct comparison of the U and H$\alpha$ images, discussed above
(Figure~4, panels a-c), shows that the morphology of the blue stars
closely follows that of the ionized gas. This suggests that the U-band
image of the center of NGC5236 is tracing the optically detectable
starburst population. Figure~14 shows the azimuthally-averaged profile
of the surface brightness of both H$\alpha$ and U-band in annuli of
increasing distance from the center. The two surface brightnesses have
similar half-light radii, around 5$^{\prime\prime}$.5, with profiles
of almost identical shape. In both cases the assumption is that the
emission of the underlying non-starburst stellar population is fairly
constant out to $\sim$40$^{\prime\prime}$ (see Figure~14). The almost
identical values of the half-light radii also confirms that the gas
emission in the center of NGC5236 is as extended as the starburst
population, and there is no evidence for `leakage' of ionized photons
beyond the starburst region.
In summary, the characteristics of the nebular emission in the
starbursting center of NGC5236 are typical of gas excited
predominantly by photoionization. This conclusion should be regarded
as preliminary, for the following three reasons: (1) the
continuum-subtracted narrow-band images of NGC5236 are less deep than
those of NGC5253, especially the crucial [OIII] image; (2) the
contribution of the [NII] lines to the H$\alpha$ emission is almost 4
times higher in NGC5236 than in NGC5253; the impact of variations
of the [NII]/H$\alpha$ line ratio on the [SII]/H$\alpha$ map is
moderate if the [NII]/H$\alpha$ changes by less than 50\%~ but
increases for larger variations; (3) we do not have
appropriate information to correct for variations of the dust
reddening across the central region. Our [OIII] maps of NGC5236 contain
limited information despite the long exposure times (see
Figure~4f). This is a consequence of the large metallicity in this
galaxy: higher metallicities correspond to lower intensities for the
O$^{++}$ ion emission.
Incidentally, while [SII]/H$\alpha$ has typical values in the range
0.15--0.4 in the Northern spiral arm of the galaxy, the ratio
covers the range 0.2--1.4 and has a more extended cross section in the
Southern spiral arm (Figure~2c). A difference between the two arms is
evident neither from the stellar continuum morphology, nor from the
colors. Since we cannot easily discriminate between dust reddening and
age of the stellar population, a difference between the intrinsic
stellar populations in the two arms cannot be excluded.
\section{Discussion}
The investigation of the morphology and physical conditions of the
ionized gas in the galaxy pair NGC5253/NGC5236 demonstrates that the
ionization structure is different in the two central starbursts, and
probably reflects the morphological difference of the host
galaxies. Both galaxies responded to a possibly common trigger with a
large scale central starburst (size of order 500~pc), but the
starburst in NGC5253 was probably more extended in the past (Caldwell
\& Phillips 1989). NGC5236 is also experiencing a somewhat milder
event in its center, with a SFR about 1/3--1/4 that of NGC5253,
although dust reddening corrections are uncertain. The major
difference between the two starbursts appears to be, however, in their
impact on the surrounding ISM, as discussed below.
\subsection{NGC5253}
\subsubsection{The Diffuse Ionized Medium}
The central concentration of the blue stars relative to the ionized
gas seen in NGC5253 is typical of intense star-forming events, ranging
from giant HII regions like 30~Dor or NGC604 (Kennicutt \& Chu 1994,
Mu\~noz-Tu\~non 1994), to Blue Compact Dwarf galaxies (e.g., Meurer et
al. 1992). The structure of the starburst in NGC5253 closely matches
these expectations. The extended ionized gas emission is probably the
manifestation of the hydrodynamic effects of radiation pressure and
stellar winds/supernovae on the ISM surrounding the starburst, in
combination with the photoionization effects of luminous sources.
This conceptual model accounts for complex structures, like bubbles and
filaments, in the low surface brightness H$\alpha$ emission. In terms
of luminosity, the H$\alpha$ flux which is not directly associated
with massive stars and, thus, can be associated with the DIG is 13\%~
of the total. This fraction is for the projected emission only; we do
not attempt to extrapolate it to the entire 3-dimensional distribution
as this would require ``guessing'' the gas distribution along the line
of sight. The locus of massive stars is defined as the region where the
U-band emission from the starburst is detected (Figure~3a). Thus about
one-tenth, and probably more, of the H$\alpha$ emission in NGC5253 is
spatially separated from the source of ionization.
\subsubsection{The Contribution of Shocks and Other Processes}
The peripheral regions of the ionized emission in NGC5253 show the
presence of a shock or other non-photoionization component in the gas
excitation mechanism. Although the candidate shock structures we
identify in the previous section (Figure~3h) need spectroscopic
confirmation, it is clear that some contribution to the nebular line
emission from non-photoionization (`shock') excitation needs to be
present to explain the observed line ratios.
The morphology of the `shocked' gas (Figure~3h) closely follows that
of the bubbles South-West of the starburst center and the filaments in
the Northern region. In particular, the bulk of the `shocked' area is
located at the position of the major-axis bubble, and extends in the
direction of the finger of soft X-ray emission in the map of Martin \&
Kennicutt (1995) and along the axis of the X-ray emission detected by
Strickland \& Stevens (1999). The X-ray finger of Martin \& Kennicutt
extends for about 2$^{\prime}$ away from the galaxy center, not very
different from the 1$^{\prime}$.3 scale of the `shocked' region. A
possibility therefore exists that the hot gas in overlapping
superbubbles, which is most likely responsible for the extended X-ray
emission (Strickland \& Stevens 1999), also affects the optical
emission line spectrum. Given the relatively low photon luminosity of
the X-ray sources in NGC~5253
($\leq$10$^{44}$~photons~cm$^{-2}$~s${-1}$), the X-rays should make at
modest contribution to the level of photoionization in most of
NGC~5253. However, in addition to shocks within the hot bubbles (see
Martin \& Kennicutt 1995), there is the possibility that transition
layers of warm, rapidly cooling gas exist within or on the boundaries
of these regions. One example of this type of emission region are
`turbulent mixing layers', such as those described by Slavin, Shull,
\& Begelman (1993), which could become significant sources of emission
in regions with low gas column densities. Emission line ratios from
mixing layers can mimic shocks in the diagnostic emission line ratios
which we have available, and this possibility therefore merits future
examination. However, since hot ejecta from the central starburst can
be responsible for shocking the outer regions of the DIG, we discuss
below the viability of shocks to explain the observed line ratios.
A size scale of about 1$^{\prime}$.3 for the shocked region
corresponds to a physical size of 1.45~kpc, which, for an expansion
velocity of 35~km~s$^{-1}$ (Marlowe et al. 1995), corresponds to an
age of 40~Myr for the bubble. Star formation within the central
$\sim$300~pc has been ongoing for $\sim$100~Myr, long enough to drive
such a shock (Calzetti et al. 1997). The low velocity values observed
by Marlowe et al. represent a potential difficulty for interpreting
the ionization of this region as due to shocks. However, the
line-of-sight velocity may not be representative of the expansion
velocity of the bubble; if we adopt a shock velocity of 100~km/s, the
age of the region decreases to $\sim$15~Myr.
The H$\alpha$ intensity associated with the shocked component is
2.2\%~ of the total H$\alpha$ emission, after correction for
underlying stellar absorption and dust obscuration (accounting only
the area for which [SII]/H$\alpha$ is detected above 5~$\sigma$, see
Figure~3). This value is an upper limit, as the regions we define as
`shocked' can also be partially photoionized; we assume,
conservatively, that only half, or 1.1\%, of the ionized gas emission
in the `shocked' areas is actually excited by shocks. This fraction,
however, does not include the fainter, more extended DIG, since here
[SII] is either undetected or is detected below our 5~$\sigma$
cut. Whichever the actual shocked H$\alpha$ emission fraction, it will
still be a few percent at most. The H$\alpha$ flux associated with the
shock is 3.2~E$-$13~erg~s$^{-1}$~cm$^{-2}$, corresponding to a
luminosity of 6.1$\times$10$^{38}$~erg~s$^{-1}$ at the distance of
NGC5253.
We can compare this value with the amount of mechanical energy input
expected from the starburst. The reddening-corrected flux density at
2,600~\AA~ from the star-forming region is
F(2600)=4.5~E$-$13~erg~s$^{-1}$~cm$^{-2}$~\AA$^{-1}$, corresponding to
a luminosity of $\sim$8.6$\times$10$^{38}$~erg~s$^{-1}$~\AA$^{-1}$,
from the data of Calzetti et al. (1997). This estimate does not take
into account the fraction of massive stars so deeply buried in dust,
e.g., along the dust lane, that their accounting is missing from the
UV flux; a comparison between the optical nebular emission and the
radio thermal emission (Beck et al. 1996) shows that the missing
fraction amounts to $\approx$20\%~ of the reddening-corrected UV flux.
Given the star formation history of the galaxy and the above UV flux,
the mechanical energy being deposited into the ISM by massive stars is
$\sim$8.5$\times$10$^{40}$~erg~s$^{-1}$ (Leitherer \& Heckman 1995), a
value very similar to what calculated by Marlowe et al. (1995) and by
Martin \& Kennicutt (1995). This energy rate is about 50\% higher than
the one derived for an expanding superbubble with the age and size
given above and density of 12~cm$^{-3}$ (Martin \& Kennicutt 1995),
using the self-similar solution of Weaver et al. (1977), and is more
than sufficient to produce the observed X-ray luminosity (Strickland
\& Stevens 1999). About 2.5\% of the shock input power is emitted in
H$\alpha$ (Binette, Dopita \& Tuohy 1985), implying that the gas
shocked by the starburst in NGC5253 can produce a total luminosity
L(H$\alpha$)$\sim$2.1$\times$10$^{39}$~erg~s$^{-1}$.
Most of the detected `shocked' gas component is located in the south-western
quadrant relative to a sphere centered on the main cluster. If the
mechanical energy is emitted with bi-polar symmetry from the central
starburst (Strickland \& Stevens 1999), this quadrant is likely to
receive $\approx$25\%~ of the total energy available to shocks, or
L(H$\alpha$)$\sim$5.3$\times$10$^{38}$~erg~s$^{-1}$, comparable to the
observed luminosity of 6.1$\times$10$^{38}$~erg~s$^{-1}$. The total
mechanical energy available from the starburst can shock-excite gas to
produce a total H$\alpha$ luminosity $\sim$3.4~times what observed or
about one-fourth of the 13\% of H$\alpha$ emission we associate with
the DIG. Observationally photoionization provides between 80\% and
90\% of the excitation of the DIG in NGC5253; this fraction is
potentially lower, but not lower than $\sim$70\%, even if all the
mechanical energy were available to excite the gas.
Notably absent are shocks in the Eastern region; in section~3.2.5 we
showed that the ionized gas associated with the dust lane in the E
region is more consistent with photoionization rather than shocks (or
other mechanism), even though local stellar ionizing sources have not
yet been found. The absence of an obvious shock component here has
another intepretation, possibly complementary to the previous one.
The dust lane coincides with the position of the extremely weak CO
detection in this galaxy (Turner et al. 1997). Turner et al. have
interpreted the very low CO luminosity as evidence for the presence of
extremely metal-poor gas in the area, possibly infalling gas which is
feeding the central starburst. If the metallicity along the dust lane
is lower than the average in the starburst, the [OIII]/H$\beta$ and
the [SII]/H$\alpha$ ratios are expected to be higher and lower,
respectively, than the average. Thus, presence of shocks along this
region would go undetected by our method, as we are assuming a uniform
metallicity for the ionized gas across the entire central region.
\subsubsection{The Structure of the Starburst}
Shocks in NGC5253 appear to have a preferential direction along the
galaxy's major axis. In addition the region along the dust lane,
namely along the optical minor axis, appears dominated by
photoionization. These facts, together with the HI kinematical data
of Kobulnicky \& Skillman (1995) and the CO detection of Turner et
al. (1997) suggest the following picture for the starburst in
NGC5253:
\begin{enumerate}
\item The central star formation is being fueled by gas which is
either infalling along the minor axis, as suggested by Turner et al.,
or is located in a `disk' rotating about the major axis, as suggested
by Kobulnicky \& Skillman on the basis of the HI rotation.
\item Hot ejecta from supernovae explosions and stellar winds drive
the expansion of the ISM described in Martin \& Kennicutt (1995) and
in Strickland \& Stevens (1999). The expansion is driven mostly in the
direction perpendicular to the disk/infalling~gas, where the gas
density is lower (or, alternatively, is driven along the gas disk
rotation axis). Hot ejecta from the central starburst, thus, shock
the gas preferentially along the optical major axis. Also, the shocked
gas is detected mainly along the south-western side of the major axis;
the propagation of the northward shocks may be prevented by the high
density region of the infalling-gas/rotating-gas-disk, which is
located in the northern side of the star forming site.
\end{enumerate}
This general picture is rather different from the one found for other
dwarf galaxies by Marlowe et al. (1995), where the location of bubbles
is preferentially along the optical minor axis, suggesting that the
ionized gas expands mainly in the direction perpendicular to the plane
of the galaxy. One can speculate that past interaction with NGC5236
played a role in the geometry of NGC5253; the massive `companion' is
located in the NW quadrant, at position angle P.A.$\sim-$20$^o$. This
direction is only $\sim$25$^o$ away from the minor axis and almost
orthogonal, just $\sim$15--20$^o$ away, to the direction of the
shocked gas. Thus, the direction along which the encounter between the
two galaxies happened may have determined the initial gas infall and
subsequent gas expansion directions in NGC5253.
\subsection{NGC5236}
For NGC5236, the gas morphology allows a more straightforward
interpretation than in NGC5253, although our conclusions are limited
by the shallowness of the [OIII] map and by presence of large amount
of dust in combination with the lack of an H$\beta$ image to perform
dust reddening corrections. The H$\alpha$ emission correlates fairly
well with the blue emission from the ionizing stars, and there is no
evidence for extended ionized gas emission. The strong spatial
overlap between blue stars and ionized gas indicates that we are not
seeing any `bonafide' DIG in the center of this galaxy; rather,
ionization appears to be a local process. The [SII]/H$\alpha$ values
fall into the photoionization range even after allowing for large
uncertainties in the underlying stellar absorption and in the [NII]
contribution, suggesting very little, if any, contribution from a
shock or other non-photoionization component.
The central starburst in NGC5236 is a milder perturbation on its giant
spiral galaxy host than the one in NGC5253. The past encounter with
NGC5253 may have produced a stellar bar and/or triggered the gas
inflow towards the center along the bar. The presence of an inner
Lindblad resonance (Telesco et al. 1993) is the additional ingredient
needed to produce a ring starburst (Shlosman, Begelman \& Frank
1990). The absence of a shocked component in the ionized gas can be
interpreted as an effect of the deep potential well in the center of
NGC5236. The more massive the galaxy, the harder it is to disrupt the
gas disk, especially in a high density center (De Young \& Heckman
1994, MacLow \& Ferrara 1998). In NGC5236, the mechanical energy being
deposited into the ISM by the central starburst is
$\approx$2$\times$10$^{40}$~erg~s$^{-1}$. This amount of power is
probably inadequate to disrupt the ISM in the dense nuclear disk of
NGC5236. The higher gas densities limit the growth of superbubbles,
while the larger gravitational forces make expansion out of the plane
more difficult. The calculations of De Young \& Heckman refer to
disruption along the minor axis, while we are looking at ISM expansion
parallel to the gas disk (NGC5236 is seen nearly face-on). However, if
the ISM is left intact along the minor axis, it is even more likely to
remain confined in the center of the disk. Lack of obvious filamentary
structures, bubbles and superbubbles in the ionized gas of NGC5236
fits in this picture.
\section{Conclusions}
The analysis of the starburst galaxy `odd couple' NGC5253 and NGC5236
reveals very different ionized gas morphologies. The metal-poor, dwarf
member of the pair, NGC5253, has the DIG emission typical of intense
bursts of star formation, that accounts for about 13\% of the
projected H$\alpha$ luminosity. A small ($\sim$10--20\%), but not
negligible, fraction of the DIG is ionized by shocks or other
non-photoionization mechanism; this implies that between 80\% and 90\%
of the H$\alpha$ emission from the DIG is due to photoionization from
massive stars. The morphology of the `shocked' gas is quite peculiar,
as it extends along the optical major axis, orthogonal to the
direction from which presumably the gas is feeding the central
starburst. If the `shocked' gas corresponds to one or more expanding
bubbles driven by the central starburst, the in-plane morphology
indicates that the metals ejected from the central region will remain
inside the galaxy, and will not be lost in the intergalactic medium
(Mac Low \& Ferrara 1998). Photoionization of the DIG from massive
stars means that about 10\%~ of the ionizing photons are escaping from
the central starburst zone.
In the metal-rich, grand-design spiral member of the pair, NGC5236,
there is no clear detection of a DIG component in the starbursting
nuclear region and the ionized gas does not show an obvious shocked
component. This is probably because the gas is confined to the center
by the deep potential well of the galaxy, and remains near the
massive stars responsible for its photoionization.
The fraction of DIG to total ionized gas in both starbursts is much
smaller, probably due to projection effects, than the 20--50\%
measured in less active star-forming galaxies (Ferguson et al. 1996a,
Wang et al. 1997, 1998). Although we can only place a lower limit to
the amount of DIG in the two starbursts, it is unlikely that the actual
fraction will be higher than what has been observed in other galaxies.
\acknowledgments
D.C., C.J.C., and A.L.K. thank the Carnegie Observatories for the
hospitality and for granting them observing time at the Las Campanas
Observatory. D.C. thanks Crystal Martin for useful discussions and
suggestions during the analysis of the images. Part of this manuscript
was written at the Kitt Peak 2.1-m telescope, during a stormy night.
\clearpage
\begin{deluxetable}{lrrrr}
\large
\tablecaption{Summary of Filters, Calibrations, Limiting
Sensitivities. \label{tbl-1}}
\tablewidth{0pt}
\tablehead{
\colhead{Filter} & \colhead{$\lambda_c$\tablenotemark{a}} & \colhead{FWHM} & \colhead{Flux Conversion\tablenotemark{b}} & Detection Limit\tablenotemark{c}\\
\colhead{ } & \colhead{(\AA)} & \colhead{(\AA)} &
\colhead{(erg cm$^{-2}$ \AA$^{-1}$ ADU$^{-1}$)} &
}
\startdata
Harris U & & & 2.139~E$-$18 (5\%) & 4.6~E$-$20 \nl
Harris V & & & 2.550~E$-$19 (2\%) & 4.3~E$-$20 \nl
Harris R & & & 1.160~E$-$19 (1.5\%) & 3.6~E$-$20\nl
5000/70 & 4994 & 77 & 4.697~E$-$18 (4\%) & 7.4~E$-$18 \nl
6563/78 & 6568 & 68 & 2.361~E$-$18 (3\%) & 8.0~E$-$18 \nl
6737/76 & 6747 & 91 & 1.796~E$-$18 (3\%) & 9.3~E$-$18 \nl
\enddata
\tablenotetext{a}{The central wavelength of the narrow band filters.}
\tablenotetext{b}{The flux zeropoint is given with, in parenthesis,
the internal uncertainty.}
\tablenotetext{c}{The limiting surface brightness is in
erg~s~cm$^{-2}$~arcsec$^{-2}$ for the continuum-subtracted narrow band
images ([OIII], H$\alpha$, and [SII]) and it is a surface brightness
density in erg~s~cm$^{-2}$~arcsec$^{-2}$~\AA$^{-1}$ for the broad band
images. The values refer to 1~$\sigma$ detection limits of the deepest
images obtained in this project (see Table~2), rebinned to a
resolution of 1$^{\prime\prime}$.3, namely 5$\times$5~pix$^2$.}
\end{deluxetable}
\clearpage
\begin{deluxetable}{lrrrr}
\large
\tablecaption{Summary of Exposure Times. \label{tbl-2}}
\tablewidth{0pt}
\tablehead{
\colhead{ } & \colhead{NGC5253} & \colhead{NGC5253} & \colhead{NGC5236}
& \colhead{NGC5236}\\
\colhead{Filter} & \colhead{Exp. Time\tablenotemark{a}} & \colhead{Exp. Time\tablenotemark{a}} & \colhead{Exp. Time\tablenotemark{a}} & \colhead{Exp. Time\tablenotemark{a}}\\
\colhead{ } & \colhead{(s)} & \colhead{(s)} & \colhead{(s)} & \colhead{(s)}
}
\startdata
Harris U & 900. & 8400. & 210. & 4500. \nl
Harris V & 120. & 1200. & 190. & 920. \nl
Harris R & 240. & 1520. & 260. & \nodata \nl
5000/70 & 180. & 10200. & 870.& 3000. \nl
6563/78 & 240. & 3900. & 780. & 1440. \nl
6737/76 & 12000. & \nodata & 1560. & 3840. \nl
\enddata
\tablenotetext{a}{The second and fourth columns are the total exposure
times of the final unsaturated image for NGC5253 and NGC5236,
respectively; the third and fifth columns are the total exposure times
of the final images which have the central
$\sim$5$^{\prime\prime}$--20$^{\prime\prime}$ of the galaxies
saturated. Each unsaturated/saturated image is the combination of
multiple exposures in the range 30--1200~s.}
\end{deluxetable}
\clearpage
|
\section{Introduction}
\label{sec:intro}
\par
The hybrid inflationary scenario \cite{hybrid}, which can
reproduce the measurements of the cosmic background explorer
(COBE) \cite{cobe} with more or less `natural'
values of the relevant coupling constants, is almost
automatically realized \cite{lyth,dss} in supersymmetric
grand unified theories (GUTs). In particular,
a moderate extension of the minimal supersymmetric standard
model (MSSM) based on a left-right symmetric gauge group
provides \cite{lss} a `natural' framework for the
implementation of hybrid inflation. The $\mu$ problem of
MSSM can be easily resolved \cite{dls} in the context
of this model by coupling the inflaton system to the
electroweak higgs superfields.
\par
At the end of inflation, the inflaton (oscillating system)
predominantly decays into electroweak higgs superfields,
thereby `reheating' the universe. However, its subdominant
decay mode to right handed neutrinos leads \cite{atmo},
via their subsequent decay, to the production of a
primordial lepton asymmetry in the universe. Nonperturbative
electroweak sphaleron effects, which violate baryon and
lepton number, then partially transform this asymmetry to
the observed baryon asymmetry of the universe (BAU).
\par
We analyze the consequences of this baryogenesis
mechanism on $\nu_{\mu}-\nu_{\tau}$ mixing.
We find \cite{atmo} that, for masses of $\nu_{\mu}$,
$\nu_{\tau}$ which are consistent with the small angle
MSW resolution of the solar neutrino problem and the
recent results of the SuperKamiokande experiment
\cite{superk}, maximal $\nu_{\mu}-\nu_{\tau}$
mixing can be achieved. The required values of the
relevant parameters are, however, quite small.
\par
In Sec.\ref{sec:hybrid}, we review the nonsupersymmetric
(Sec.\ref{subsec:nonsusy}) as well as the supersymmetric
(Sec.\ref{subsec:susy}) version of the hybrid
inflationary scenario. In Sec.\ref{sec:baryons}, we discuss
baryogenesis through a primordial leptogenesis. In
particular, Sec.\ref{subsec:leptons} is devoted to the
generation of the primordial lepton number. The
topologically nontrivial structure of the vacuum in gauge
theories and the resulting nonperturbative baryon and
lepton number violating phenomena in the standard model
are analyzed in Sec.\ref{subsec:sphaleron}. The rate of
these phenomena at finite temperatures is calculated by
employing electroweak sphalerons and the final BAU is
estimated. Finally, in Sec.\ref{sec:lr} the supersymmetric
model based on a left-right symmetric gauge group is
presented. In particular, the solution of the $\mu$ problem
(Sec.\ref{subsec:mu}), inflation (Sec.\ref{subsec:inf})
and leptogenesis (Sec.\ref{subsec:reheatlepto}) are sketched.
\section{Hybrid Inflation}
\label{sec:hybrid}
\subsection{The non Supersymmetric Version}
\label{subsec:nonsusy}
\par
The most important disadvantage of most inflationary
scenarios was that they needed extremely small coupling
constants in order to reproduce the results of
COBE \cite{cobe}. This difficulty
was overcome some years ago by Linde \cite{hybrid} who
proposed, in the context of nonsupersymmetric GUTs,
a clever inflationary scenario known as
hybrid inflation. The idea was to use two real scalar fields
$\chi$ and $\sigma$ instead of one that was normally used.
The field $\chi$ provides the vacuum energy which drives
inflation while $\sigma$ is the slowly varying field
during inflation. The main advantage of this scenario is
that it can reproduce the observed temperature fluctuations
of the cosmic background radiation (CBR) with `natural'
values of the parameters in contrast to previous
realizations of inflation (like the `new' \cite{new} or
`chaotic' \cite{chaotic} inflationary scenarios). The
potential utilized by Linde is
\begin{equation}
V ( \chi, \sigma)= \kappa^2 \left( M^2 -
\frac {\chi^2}{4}\right)^2 + \frac{\lambda^2
\chi^2 \sigma^2} {4} + \frac {m^2\sigma^2}{2}~~,
\label{eq:lindepot}
\end{equation}
where $\kappa,~\lambda$ are dimensionless positive
coupling constants and $M$, $m$ mass parameters. The
vacua lie at $\langle \chi\rangle=\pm 2 M$, $\langle
\sigma \rangle=0$. Putting $m$=0, for the moment, we
observe that the potential possesses an exactly flat
direction at $\chi=0$ with $V(\chi=0 ,\sigma)=\kappa^2
M^4$. The mass squared of the field $\chi$ along this flat
direction is given by $m^2_\chi=-\kappa^2 M^2+\frac{1}{2}
\lambda^2 \sigma^2$ and remains nonnegative for $\sigma
\geq \sigma_c = \sqrt{2}\kappa M/\lambda $. This means
that, at $\chi=0$ and $\sigma \geq \sigma_c$, we obtain
a valley of minima with flat bottom. Reintroducing the
mass parameter $m$ in Eq.(\ref{eq:lindepot}), we observe
that this valley acquires a nonzero slope. A region of the
universe, where $\chi$ and $\sigma$ happen to be almost
uniform with negligible kinetic energies and with values
close to the bottom of the valley of minima, follows this
valley in its subsequent evolution and undergoes inflation.
The quadrupole anisotropy of CBR produced during this
inflation can be estimated to be
\begin{equation}
\left(\frac {\delta T}{T}\right)_{Q}
\approx \left(\frac {16 \pi}{45}\right)^{1/2}
\frac{\lambda \kappa^2 M^5}
{M^3_Pm^2}~,
\label{eq:lindetemp}
\end{equation}
where $M_{P}=1.22\times 10^{19}{\rm{GeV}}$ is the Planck
scale. The COBE~\cite{cobe} result, $(\delta T/T)_{Q}
\approx 6.6 \times 10^{-6}$, can then be reproduced with
$M\approx 2.86\times 10^{16}$ GeV, the supersymmetric GUT
vacuum expectation value (vev), and $m \approx 1.3~\kappa
\sqrt {\lambda}\times 10^{15}$ GeV $\sim 10^{12}$ GeV
for $\kappa, \lambda \sim 10^{-2}$. Inflation terminates
abruptly at $\sigma=\sigma_{c}$ and is followed by a
`waterfall', i.e., a sudden entrance into an oscillatory
phase about a global minimum. Since the system can fall into
either of the two available global minima with equal
probability, topological defects are copiously produced if
they are predicted by the particular particle physics model
one is considering.
\subsection{The Supersymmetric Version}
\label{subsec:susy}
\par
The hybrid inflationary scenario is \cite{lyth} `tailor
made' for application to supersymmetric GUTs except that
the mass of $\sigma$, $m$, is unacceptably large for
supersymmetry, where all scalar fields acquire masses of
order $m_{3/2} \sim 1$ TeV (the gravitino mass) from
soft supersymmetry breaking. To see this, consider a
supersymmetric GUT with a (semi-simple) gauge group $G$ of
rank $\geq 5$ with $G \to G_S$ (the standard model gauge
group) at a scale $M \sim 10^{16}$ GeV. The spectrum of
the theory below $M$ is assumed to coincide with the
MSSM spectrum plus standard
model singlets so that the successful predictions for
$\alpha_{s}$, ${\rm{sin}}^{2} \theta_{W}$ are retained.
The theory may also possess global symmetries. The breaking
of $G$ is achieved through the superpotential
\begin {equation}
W =\kappa S( \phi\bar{\phi}- M^2),
\label{eq:superpot}
\end {equation}
where $\phi$, $\bar{\phi}$ is a conjugate pair of
standard model singlet left handed superfields which belong
to nontrivial representations of $G$ and reduce its rank
by their vevs and $S$ is a gauge singlet left handed
superfield. The coupling constant $\kappa$ and the mass
parameter $M$ can be made positive by phase redefinitions.
This superpotential is the most general renormalizable
superpotential consistent with a $U(1)$ R-symmetry under
which $W \to e^{i\theta} W,~S \to e^{i \theta}S,~\phi
\bar {\phi} \to \phi\bar{\phi}$ and gives the potential
\begin{equation}
V=\kappa^2 \mid M^2-\phi\bar{\phi}\mid^2
+\kappa^2 \mid S \mid^2
(\mid \phi \mid^2 +\mid \bar{\phi}\mid^2)
+{\rm{ D-terms}}.
\label{eq:hybpot}
\end{equation}
Restricting ourselves to the D flat direction $\phi=
\bar{\phi}^* $ which contains the supersymmetric minima and
performing appropriate gauge and R- transformations, we can
bring $S$, $\phi$, $\bar{\phi}$ on the real axis, i.e.,
$S \equiv \sigma/\sqrt{2}$, $\phi=\bar{\phi} \equiv
\chi/2$, where $\sigma$, $\chi$ are normalized real scalar
fields. The potential then takes the form in
Eq.(\ref{eq:lindepot}) with $\kappa = \lambda$ and $m=0$ and,
thus, Linde's potential for hybrid inflation is almost
obtainable from supersymmetric GUTs but without the mass
term of $\sigma$ which is, however, of crucial importance
since it provides the slope of the valley of minima
necessary for inflation.
\par
One way to obtain a valley of minima useful for inflation
is \cite{lp} to replace the trilinear term
in $W$ in Eq.(\ref{eq:superpot}) by
the next order nonrenormalizable coupling. Another way,
which we will adopt here, is \cite{dss} to keep the
renormalizable superpotential in Eq.(\ref{eq:superpot})
and use the radiative corrections along the inflationary
valley ($\phi=\bar{\phi}= 0$~, $S > S_{c}\equiv M$).
In fact, due to the mass splitting in the supermultiplets
$\phi$, $\bar{\phi}$ caused by the supersymmetry
breaking `vacuum' energy density $\kappa^{2} M^{4}$
along this valley, there are important radiative corrections.
At one-loop, the inflationary potential is given
\cite{dss,lss} by
$$
V_{\rm{eff}}(S) = \kappa^{2}M^{4}
$$
\begin{equation}
\left[
1 +\frac{\kappa^2}{32 \pi^2} \left(2\ln\left(
\frac{\kappa^{2}S^{2}}{\Lambda^2}\right)+
\left(\frac{S^2}{S_{c}^2}-1\right)^2
\ln\left(1-\frac{S_{c}^2}{S^2}\right)
+\left(\frac{S^2}{S_{c}^2}+1\right)^2
\ln\left(1+\frac{S_{c}^2}{S^2}\right)\right)
\right]~,
\label{eq:veffexact}
\end{equation}
where $\Lambda$ is a suitable mass renormalization scale.
For $S$ sufficiently larger than $S_{c}$,
\begin{equation}
V_{{\rm{eff}}} (S) = \kappa^2 M^4
\left[ 1 + \frac{\kappa^2}{16\pi^2} \left( \ln
\left(\frac{\kappa^{2} S^{2}}{\Lambda^2}\right) +
\frac{3}{2} - \frac{S_c^4}{12S^4} +
\cdots \right)\right]~.
\label{eq:veff}
\end{equation}
Using this effective potential, one finds
that the cosmic microwave quadrupole anisotropy
\begin{equation}
\left(\frac{\delta T}{T}\right)_{Q} \approx 8 \pi
\left(\frac{N_{Q}}{45}\right)^{1/2}
\frac{x_{Q}}{y_{Q}}\left(\frac{M}{M_{P}}\right)^2~.
\label{eq:cobe}
\end{equation}
Here, $N_{Q}$ is the number of e-foldings experienced by
the universe between the time the quadrupole scale exited
the inflationary horizon and the end of inflation and
$y_{Q}=x_{Q}(1-7/(12x_{Q}^2)+\cdots)$ with
$x_{Q}=S_{Q}/M$, $S_{Q}$ being the value of the scalar
field $S$ when the scale which evolved to the present
horizon size crossed outside the de Sitter (inflationary)
horizon. Also, from
Eq.(\ref{eq:veff}), one finds
\begin{equation}
\kappa \approx \frac{8\pi^{3/2}}{\sqrt{N_{Q}}}
~ y_{Q}~\frac{M}{M_{P}}~.
\label{eq:kappa}
\end{equation}
\par
The inflationary phase ends as $S$ approaches $S_{c}$ from
above. Writing $S=xS_{c}$, $x=1$ corresponds to the phase
transition from $G$ to $G_{S}$ which, as it turns out,
more or less coincides with the end of the inflationary
phase as one deduces from the slow roll conditions
\cite{dss,lazarides}. Indeed, the $50-60$ e-foldings needed
for the inflationary scenario can be realized even with small
values of $x_{Q}$. For definiteness, we will take
$x_{Q}\approx 2$. From COBE \cite{cobe} one then obtains
$M \approx 5.5 \times 10^{15}~{\rm{GeV}}$ and
$\kappa\approx 4.5\times 10^{-3}$ for $N_{Q}\approx 56$.
Moreover, the primordial density fluctuation spectral index
$n \simeq 0.98$. We see that the relevant part of inflation
takes place at $S\sim 10^{16}~{\rm{GeV}}$. An interesting
consequence of this is \cite{lyth,lss,sugra} that the
supergravity corrections can be negligible.
\par
In conclusion, it is important to note that the superpotential
$W$ in Eq.(\ref{eq:superpot}) leads to hybrid inflation in a
`natural' way. This means that a) there is no need of very
small coupling constants, b) $W$ is the most general
renormalizable superpotential allowed by the gauge and R-
symmetries, c) supersymmetry guarantees that the radiative
corrections do not invalidate inflation, but rather provide
a slope along the inflationary trajectory which drives the
inflaton towards the supersymmetric vacua, and
d) supergravity corrections can be brought under control
so as to leave inflation intact.
\section{Baryogenesis via Leptogenesis}
\label{sec:baryons}
\subsection{Primordial Leptogenesis}
\label{subsec:leptons}
\par
In most hybrid inflationary models, it is not convenient to
produce the observed BAU in the customary way, i.e., through
the decay of color $~3,~\bar{3}~$ fields $(g,~g^{c})$.
Some of the reasons are the following: i) For theories where
leptons and quarks belong to different representations of
the unifying gauge group $G$ (which is the case, for example,
for $G=G_{LR}\equiv SU(3)_c\times SU(2)_L\times SU(2)_R
\times U(1)_{B-L}$ or $SU(3)_c \times SU(3)_L \times
SU(3)_R$), the baryon number can be made almost
exactly conserved by imposing an appropriate discrete
symmetry. In particular, for $G=G_{LR}$, we can impose
\cite{baryonparity} a discrete symmetry under which
$q\rightarrow - q$, $q^{c} \rightarrow - q^{c}$,
$\bar{q} \rightarrow - \bar{q}$, $\bar{q}^{c}
\rightarrow - \bar{q}^{c}$ and all other superfields
remain invariant ($q,~q^{c},~\bar{q},~\bar{q}^{c}$
are superfields with the quantum numbers of the quarks,
antiquarks and their conjugates respectively). ii) For
theories where such a discrete symmetry is absent, we
could, in principle, use as inflaton a pair of conjugate
standard model singlet superfields $N$, $\bar{N}$ which
decay into $g$, $g^{c}$. For $G=SU(3)_c \times SU(3)_L
\times SU(3)_R$, $N$ ($\bar{N}$) could be the standard
model singlet component of the (1, $\bar{3}$, 3)
( (1, 3, $\bar{3}$) ) superfields with zero $U(1)_{B-L}$
charge. But this is again unacceptable since the breaking
of $SU(3)_c \times SU(3)_L \times SU(3)_R$ by the vevs
of $N$, $\bar{N}$ predicts \cite{trinifiedmonopoles}
magnetic monopoles which can then be copiously produced
after inflation. Also the gravitino constraint
\cite{gravitino} on the `reheat' temperature, $T_{r}
\stackrel{_{<}}{_{\sim }} 10^9$ GeV, implies $m_g
\stackrel{_{<}}{_{\sim }} 10^{10}$ GeV (~from the
coupling ($~m_g / \langle N\rangle) Ngg^{c}$~)
leading to strong deviation from MSSM and possibly
proton decay.
\par
So it is preferable to produce first a primordial lepton
asymmetry \cite{lepto} which can then be partially turned
into the observed baryon asymmetry of the universe by the
nonperturbative sphaleron effects \cite{sphaleron} of
the electroweak sector. In the particular model based on
$G_{LR}$ which we will consider later, this is the only
way to produce the BAU since the inflaton decays into
higgs superfields and right handed neutrinos. The subsequent
decay of right handed neutrinos into ordinary higgs
particles (higgsinos) and light leptons (sleptons) can
produce the primordial lepton asymmetry. It is important,
though, to ensure that this primordial lepton asymmetry
is not erased \cite{turner} by lepton number violating
$2 \rightarrow 2$ scattering processes such as
$ l l \rightarrow h^{(1)}\,^{*} h^{(1)}\,^{*}$ or
$l h^{(1)} \rightarrow \bar{l} h^{(1)}\,^{*}$ ($l$
represents a lepton doublet and $h^{(1)}$ the higgs
superfield which couples to the up type quarks) at all
temperatures between $T_{r}$ and 100 GeV. This is
automatically satisfied since the primordial lepton
asymmetry is protected \cite{ibanez} by supersymmetry
at temperatures between $T_{r}$ and $T \sim 10^{7}$ GeV,
and for $T \stackrel{_{<}}{_{\sim }} 10^{7}$ GeV,
these $2 \rightarrow 2$ scattering processes are well
out of equilibrium provided \cite{ibanez}
$m_{\nu_{\tau}}\stackrel{_<}{_\sim} 10~{\rm{eV}}$,
which readily holds in our case (see below).
\par
The lepton asymmetry produced by the out-of-equilibrium
decay ($M_{\nu^{c}_{i}} \gg T_{r})$ of the right
handed neutrinos $\nu^{c}_{i}$, which emerged from the
inflaton decay, is \cite{lepto}
\begin{equation}
\frac {n_{L}}{s} \approx - \frac{3}{16 \pi}
\frac {T_{r}}{m_{infl}}\sum_{l \neq i}
g(r_{li}) \frac{{\rm{Im}}(U~M^{D}\,^{\prime}
~M^{D}\,^{\prime}\,^{\dagger}~U ^{\dagger})^{2}_{il}}
{|\langle h^{(1)}\rangle|^{2}(U~M^{D}\,^{\prime}
~M^{D}\,^{\prime}\,^{\dagger}~U ^{\dagger})_{ii}}~,
\label{eq:genlept}
\end{equation}
where $n_L$ and $s$ are the lepton number and entropy
densities, $m_{infl}$ in the inflaton mass,
$M^{D}\,^{\prime}$ is the diagonal `Dirac'
mass matrix, $U$ a unitary transformation so that
$UM^{D}\,^{\prime}$ is the `Dirac' mass matrix in the
basis where the `Majorana' mass matrix of $\nu^{c}$~'s
is diagonal and $|\langle h^{(1)} \rangle| \approx
174~{\rm{GeV}}$ for large ${\rm tan}\beta$. The function
\begin{equation}
g(r_{li}) = r_{li}~\ln \left(\frac {1 + r^{2}_{li}}
{r^{2}_{li}}\right)~,~r_{li} = \frac {M_{l}} {M_{i}}~,
\label{eq:gfunction}
\end{equation}
with $g(r) \sim 1/r$ as $r \rightarrow \infty$.
Here we have taken into account the following prefactors:
i) At `reheat', $n_{infl} m_{infl} =
(\pi^{2}/30) g_{*} T^{4}_{r}$ ($n_{infl}$ is the
inflaton number density and $g_{*}$ the effective
number of massless degrees of freedom) which together
with the relation $s = (2 \pi^{2}/45) g_{*}T^{3}_{r}$
implies that $n_{infl}/s = (3/4)(T_{r}/m_{infl})$.
ii) Since each inflaton decays into two $\nu^{c}$~'s,
their number density $n_{\nu^{c}} = 2 n_{infl}$ which
then gives $n_{\nu^{c}}/s= (3/2)(T_{r}/m_{infl})$.
iii) Supersymmetry gives an extra factor of two.
\subsection {Sphaleron Effects}
\label{subsec:sphaleron}
\par
To see how the primordial lepton asymmetry partially turns
into the observed BAU, we must first discuss the
nonperturbative baryon ($B$-) and lepton ($L$-) number
violation \cite{thooft} in the standard model. Consider the
electroweak gauge symmetry $SU(2)_{L} \times U(1)_{Y}$
in the limit where the Weinberg angle $\theta_W=0$ and
concentrate on $SU(2)_{L}$ (inclusion of $\theta_{W}\neq 0$
does not alter the conclusions). Also, for the moment, ignore
the fermions and higgs fields so as to have a pure
$SU(2)_{L}$ gauge theory. This theory has \cite{vacuum}
infinitely many classical vacua which are topologically
distinct and are characterized by a `winding number'
$n \in Z$. In the `temporal gauge' ($A_{0}=0$), the
remaining gauge freedom consists of time independent
transformations and the vacuum corresponds to a pure gauge
\begin{equation}
A_{i} = \frac{i}{g}~\partial _{i}g(\bar{x})
g ^{-1}(\bar{x})~,
\label{eq:gauge}
\end{equation}
where $g$ is the $SU(2)_{L}$ gauge coupling constant,
$\bar{x}$ belongs to ordinary 3-space, $i$ =1,2,3,
$g(\bar{x})
\in SU(2)_{L}$, and $g(\bar{x}) \rightarrow 1 $ as
$ \mid\bar{x}\mid \rightarrow \infty$. Thus, the 3-space
compactifies to a sphere $S^{3}$ and $g(\bar{x})$ defines
a map: $S^{3} \rightarrow SU(2)_{L}$ (with the
$SU(2)_{L}$ group being topologically equivalent to
$S^{3}$). These maps are classified into homotopy classes
constituting the third homotopy group of $S^{3},~\pi_{3}
(S^{3})$, and are characterized by a `winding number'
\begin{equation}
n = \int d^{3}x~\epsilon^{ijk}~{\rm{tr}}
\left(\partial_{i}g(\bar{x})
g^{-1}(\bar{x})\partial_{j}g(\bar{x}) g^{-1}(\bar{x})
\partial_{k}g(\bar{x}) g^{-1}(\bar{x})\right).
\label{eq:wind}
\end{equation}
The corresponding vacua are denoted as $\mid n\rangle$,
$n\in Z$.
\par
The tunneling amplitude from the vacuum $\mid n_{-}
\rangle$ at $t=-\infty$ to the vacuum $\mid n_{+}
\rangle~$ at $t=+\infty$ is given by the functional
integral
\begin{equation}
\langle n_{+}\mid n_{-}\rangle = \int(dA)~e^{-S(A)}
\label{eq:path}
\end{equation}
over all gauge field configurations satisfying the
appropriate boundary conditions at $t=\pm \infty$.
Performing a Wick rotation,
~$x_0 \equiv t \rightarrow -i x_{4}$,
we can go to Euclidean space-time. Any Euclidean field
configuration with finite action is characterized by an
integer topological number known as the Pontryagin number
\begin{equation}
q = \frac {g^{2}}{16\pi^{2}}
\int d^{4}x~{\rm{tr}}\left(F^{\mu \nu}
\tilde{F}_{ \mu \nu}\right)~,
\label{eq:pontryagin}
\end{equation}
with $\mu$,$\nu$=1,2,3,4 and $\tilde{F}_{\mu \nu}=
\frac {1}{2}\epsilon_{\mu \nu \lambda \rho}
F^{\lambda \rho}$ being the dual field strength. But
${\rm{tr}} (F^{\mu \nu} \tilde{F}_{\mu \nu}) =
\partial ^{\mu} J_{\mu}$, where $J_{\mu}$ is the
`Chern-Simons current' given by
\begin{equation}
J_{\mu}=\epsilon_{\mu \nu \alpha \beta}~{\rm{tr}}
\left(A^{\nu}
F^{\alpha \beta}-\frac{2}{3}gA^{\nu}A^{\alpha}
A^{\beta}\right).
\label{eq:csc}
\end{equation}
In the `temporal gauge' ($A_0=0$),
\begin{eqnarray*}
q=\frac{g^{2}}{16 \pi^{2}} \int d^{4}x
~\partial^{\mu}J_{\mu}=
\frac{g^{2}}{16\pi^{2}}
\mathop{\Delta}_{x_{4}=\pm \infty}
\int d^{3}x~J_{0}
\end{eqnarray*}
\begin{equation}
=\frac{1}{24\pi^{2}}\mathop{\Delta}_{x_{4}=
\pm \infty}
\int d^{3}x~\epsilon^{ijk}~
{\rm{tr}}\left(\partial_{i} g g^{-1}\partial_{j}
gg^{-1}\partial_{k}gg^{-1}\right)=n_{+}-n_{-}~.
\label{eq:interpol}
\end{equation}
This means that the Euclidean field configurations which
interpolate between the vacua $\mid n_{+}\rangle,
~\mid n_{-}\rangle$ at $x_4 =\pm \infty$ have
Pontryagin number $q= n_{+}-n_{-}$ and the path integral
in Eq.(\ref{eq:path}) should be performed over all these
field configurations.
\par
For a given $q$, there is a lower bound on $S(A)$,
\begin{equation}
S(A) \geq \frac{8 \pi^{2}}{g^{2}}\mid q \mid~,
\label{eq:lbound}
\end{equation}
which is saturated if and only if $F_{\mu \nu}=
\pm \tilde{F} _{\mu \nu}$, i.e, if the
configuration is self-dual or self-antidual. For $q$=1,
the self-dual classical solution is called instanton
\cite{instanton} and is given by (in the `singular' gauge)
\begin{equation}
A_{a \mu}(x)=\frac{2 \rho^{2}}{g(x-z)^{2}}
~\frac{\eta_{a \mu \nu} (x-z)^{\nu}}
{(x-z)^{2} + \rho^{2}}~,
\label{eq:instanton}
\end{equation}
where $\eta_{a \mu \nu}$ ($a$=1,2,3; $\mu$,$\nu$=
1,2,3,4) are the t' Hooft symbols with $\eta_{aij}=
\epsilon_{aij}$ ($i$,$j$=1,2,3), $\eta_{a4i}=
-\delta_{ai}$, $\eta_{ai4}=\delta_{ai}$ and
$\eta_{a44}=0$. The instanton depends on four
Euclidean coordinates $z_{\mu}$ (its position) and
its scale (or size) $\rho$. Two successive vacua
$\mid n\rangle$,~$\mid n+1\rangle$ are separated by
a potential barrier of height $\propto \rho^{-1}$.
The Euclidean action of the interpolating instanton is
always equal to $8 \pi^{2}/g^{2}$, but the height of
the barrier can be made arbitrarily small since the size
$\rho$ of the instanton can be taken arbitrarily large.
\par
We now reintroduce the fermions into the theory and
observe~\cite{thooft} that the $B$- and $L$- number
currents carry anomalies, i.e.,
\begin{equation}
\partial_{\mu} J^{\mu}_{B} =
\partial _{\mu} J^{\mu}_{L} =
- n _{g} \frac {g^{2}}{16 \pi ^{2}}~{\rm{tr}}
(F_{\mu \nu} \tilde{F}^{\mu \nu})~,
\label{eq:anomaly}
\end{equation}
where $n_{g}$ is the number of generations. It is then
obvious that the tunneling from $\mid n_{-}\rangle$ to
$\mid n_{+}\rangle$ is accompanied by a change of the
$B$- and $L$- numbers, $\Delta B=\Delta L=- n_{g}q=
- n_{g} (n_{+}-n_{-})$. Note that i) $\Delta (B-L)=0$,
and ii) for $q$=1, $\Delta B=\Delta L=-3$ which means
that we have the annihilation of one lepton per family
and one quark per family and color (12-point function).
\par
We, finally, reintroduce the Weinberg-Salam higgs doublet
$h$ with its vev given by
\begin{equation}
<h> = \frac {v}{\sqrt{2}}~\left(\matrix{0 \cr 1 \cr}
\right)~,~v \approx 246 ~{\rm{GeV}}~.
\label{eq:vev}
\end{equation}
It is then easy to see that the instanton ceases to exist
as an exact solution. It is replaced by the so called
`restricted instanton' \cite{restricted} which is an
approximate solution for $\rho \ll v^{-1}$. For
$\mid x-z\mid \ll \rho$, the gauge field configuration
of the `restricted instanton' essentially coincides with
that of the instanton and the higgs field is
\begin{equation}
h(x) \approx \frac {v}{\sqrt{2}}
~\left(\frac{(x-z)^{2}}
{(x-z)^{2} + \rho^{2}}\right)^{1/2}
\left(\matrix {0 \cr 1 \cr } \right)~~.
\label{eq:restricted}
\end{equation}
For $\mid x-z \mid \gg \rho$, the gauge and higgs fields
decay to a pure gauge and the vev in Eq.(\ref{eq:vev})
respectively. The action of the `restricted instanton' is
$S_{ri}=(8 \pi^{2}/g^{2})+\pi^{2} v^{2} \rho^{2}+
\cdots$, which implies that the contribution of big size
`restricted instantons' to the path integral in
Eq.(\ref{eq:path}) is suppressed. This justifies
{\it a posteriori} the fact that we restricted ourselves
to approximate instanton solutions with $\rho \ll v^{-1}$.
\par
The height of the potential barrier between the vacua
$\mid n\rangle,~\mid n+1\rangle$ cannot be now
arbitrarily small. This can be understood by observing
that the static energy of the `restricted instanton' at
$x_{4}=z_{4}$ ($\lambda$ is the higgs self-coupling),
\begin{equation}
E_{b}(\rho) \approx \frac{3 \pi^{2}}
{g^{2}}~\frac{1}{\rho} +
\frac {3}{8}\pi^{2} v^{2} \rho^{2} +
\frac {\lambda}{4}\pi^{2} v^{4}\rho^{3}~,
\label{eq:static}
\end{equation}
is minimized for
\begin{equation}
\rho_{{\rm{min}}} = \frac {\sqrt{2}}{gv}
\left(\frac {\lambda}
{g^{2}}\right)^{-1/2}\left(\left(\frac {1}{64} +
\frac {\lambda}{g^{2}}\right)^{1/2} -
\frac {1}{8}\right)^{1/2}
\sim M^{-1}_{W}~,
\label{eq:rhomin}
\end{equation}
and, thus, the minimal height of the potential barrier
turns out to be $E_{{\rm{min}}} \sim M_{W} / \alpha_W$
($M_{W}$ is the weak mass scale and
$\alpha_{W}=g^{2}/4 \pi$).
The static solution which corresponds to the top
(saddle point) of this potential barrier is called
sphaleron \cite{sphaleronsol} and is given by
\begin{equation}
\bar{A} = v~\frac {f(\xi)}{\xi}~\hat{r}
\times \bar{\tau}~,
~h = \frac {v}{\sqrt{2}}~t(\xi)~ \hat {r}\cdot
\bar{\tau} \left (\matrix{0 \cr 1 \cr} \right),
\label{eq:sphaleron}
\end{equation}
where $\xi=2M_{W}r$, $\hat{r}$ is the radial unit vector
in ordinary 3-space
and the 3-vector $ \bar{\tau}$ consists of the Pauli
matrices. The functions $f(\xi),~t(\xi)$, which can be
determined numerically, tend to zero as $\xi \rightarrow 0$
and to 1 as $\xi \rightarrow \infty$. The mass (static
energy) of the sphaleron solution is estimated to be
\begin{equation}
E_{{\rm{sph}}} = \frac {2M_{W}}{\alpha_{W}}~k,
~1.5 \leq k \leq 2.7~,~ {\rm{for}} ~0 \leq
\lambda \leq \infty~,
\label{eq:sphmass}
\end{equation}
and lies between 10 and 15 TeV.
\par
At zero temperature the tunneling from $\mid n\rangle$ to
$\mid n+1\rangle$ is utterly suppressed \cite{thooft} by
the factor exp$(-8 \pi^{2} /g ^{2})$. At high temperatures,
however, thermal fluctuations over the potential barrier are
frequent and this transition can occur \cite{sphaleron} with
an appreciable rate. For $M_{W} \stackrel{_{<}}{_{\sim }}
T \stackrel{_{<}}{_{\sim }} T_{c}$ ($T_c$ is the
critical temperature of the electroweak transition), this
rate can be calculated \cite{sphaleron} by expanding around
the sphaleron (saddle point) solution and turns out to be
\begin{equation}
\Gamma \approx 10^{4}~ n_{g}~ \frac {v (T)^{9}}{T^{8}}
~{\rm{exp}} (-E_{{\rm{sph}}}(T)/T)~.
\label{eq:sphrate}
\end{equation}
Assuming that the electroweak phase transition is a second
order one, $v(T)$ and $E_{{\rm{sph}}}(T) \propto
(1 - T^{2}/T^{2}_{c})^{1/2}$. One can then show that
$\Gamma \gg H$ ($H$ is the Hubble parameter) for
temperatures $T$ between $\sim 200$
GeV and $\sim T_{c}$. Furthermore, for temperatures above
$T_{c}$, where the sphaleron solution ceases to exist, it
was argued~\cite{sphaleron} that we still have $\Gamma
\gg H$. The overall conclusion is that nonperturbative
$B$- and $L$- number violating processes are in
equilibrium in the universe for cosmic temperatures
$T\stackrel{_{>}}{_{\sim }} 200$ GeV. Remember that
$B-L$ is conserved by these processes.
\par
Given a primordial $L$- number density, one can
calculate \cite{turner,ibanez} the resulting $n_{B}/s$
($n_B$ is the $B$- number density). In MSSM, the
$SU(2)_{L}$ instantons produce the effective operator
(in symbolic form)
\begin{equation}
O_{2} = (q q q l)^{n_{g}} ( \tilde {h}^{(1)}
\tilde {h}^{(2)})
\tilde{W}^{4}~,
\label{eq:woperator}
\end{equation}
and the $SU(3)_{c}$ instantons the operator
\begin{equation}
O_3 = ( q q u^{c} d^{c})^{n_{g}} \tilde {g}^{6}~,
\label{eq:coperator}
\end{equation}
where $q$, $l$ are the quark, lepton $SU(2)_{L}$ doublets
respectively, $u^{c}$, $d^{c}$ the up, down type antiquark
$SU(2)_{L}$ singlets respectively, $h^{(1)}$, $h^{(2)}$
the higgses which couple to up, down type quarks respectively,
$g$, $W$ the gluons and $W$ bosons and tilde represents
their superpartners. We will assume that these interactions
together with the usual MSSM interactions are in equilibrium
at high temperatures. The equilibrium number density of
ultrarelativistic particles
$\Delta n \equiv n_{{\rm{part}}} - n_{{\rm{antipart}}}$
is given by
\begin{equation}
\frac {\Delta n} {s} = \frac {15 g}{4 \pi^{2} g_{*}}
\left (\frac {\mu}{T}\right)
\epsilon~,
\label{eq:chemical}
\end{equation}
where $g$ is the number of internal degrees of freedom of
the particle under consideration, $\mu$ its chemical
potential and $\epsilon=2$ or 1 for bosons or fermions.
For each interaction in equilibrium,
the algebraic sum of the chemical potentials of the particles
involved is zero. Solving these constraints, we end up with
only two independent chemical potentials, $\mu_{q} $ and
$\mu_{\tilde{g}}$, and the $B$- and $L$- asymmetries
are expressed \cite{ibanez} in terms of them:
$$
\frac {n_{B}}{s} = \frac {30}{4 \pi^{2}g_{*}T}
(6n_{g} \mu_{q} - (4n_{g} - 9) \mu_{\tilde{g}})~,
$$
\begin{equation}
\frac {n_{L}}{s} = - \frac {45}{4 \pi^{2} g_{*}T}
\left(\frac {n_{g}(14 n_{g} +9)}
{1+2 n_{g}} \mu_{q} + \Omega (n_{g})
\mu_{\tilde{g}}\right)~,
\label{eq:bla}
\end{equation}
where $\Omega(n_{g})$ is a known \cite{ibanez} function.
Now soft supersymmetry breaking couplings come in
equilibrium at $T \stackrel{_{<}}{_{\sim }}10^7$ GeV
since their rate $\Gamma_{S} \approx m^{2}_{3/2} /T
\stackrel{_{>}}{_{\sim }} H \approx 30~T^{2}/M_{P}$.
In particular, the nonvanishing gaugino mass implies
$\mu _{\tilde{g}} =0$ and Eqs.(\ref{eq:bla}) give
\cite{ibanez}
\begin{equation}
\frac {n_{B}}{s} = \frac {4(1+2n_{g})}{22n_{g}+13}
~\frac {n_{B-L}}{s}~.
\label{eq:bbminl}
\end{equation}
Equating $n_{B-L}/s$ with the primordial $n_{L}/s $,
we have $n_{B}/s = (- 28/79) (n_{L}/s)$, for $n_{g}=3$.
\section {The `Left-Right' Model}
\label{sec:lr}
\par
We will now study in detail a moderate extension of
MSSM based on the left-right symmetric gauge group $G_{LR}$
which provides \cite{lss} a suitable framework for hybrid
inflation. The inflaton is associated with the breaking of
$SU(2)_{R}$ and consists of a gauge singlet and a pair of
$SU(2)_{R}$ doublets. The $\mu$ problem is resolved
\cite{dls} by introducing \cite{lss,dls} a trilinear
superpotential coupling of the gauge singlet inflaton
to the electroweak higgs doublets. In the presence of
gravity-mediated supersymmetry breaking, this gauge singet
acquires a vev and, thus, generates \cite{dls},
via its coupling to the higgses, the $\mu$
term.
\par
The inflaton system, after the end of inflation,
predominantly decays into higgs superfields
and `reheats' the universe. Moreover, its subdominant
decay into right handed neutrinos provides \cite{atmo}
a mechanism for baryogenesis via leptogenesis. For
$\nu_{\mu}$, $\nu_{\tau}$ masses from the small angle
MSW resolution of the solar neutrino puzzle and the
recent results of the SuperKamiokande experiment
\cite{superk}, maximal $\nu_{\mu}-\nu_{\tau}$ mixing
can be achieved \cite{atmo}.
\subsection {The $\mu$ Problem}
\label{subsec:mu}
\par
The breaking of $SU(2)_R\times U(1)_{B-L}$ is achieved
by the renormalizable superpotential
\begin{equation}
W = \kappa S(l^c\bar l^{c}- M^2)~,
\label{W}
\end{equation}
where $S$ is a gauge singlet chiral superfield and
$l^c$, $\bar l^{c}$ is a conjugate pair of $SU(2)_R$
doublet chiral superfields which acquire superheavy
vevs of magnitude $M$. The parameters $\kappa$ and $M$
can be made positive by phase redefinitions.
\par
The $\mu$ problem can be resolved \cite{dls} by
introducing the extra superpotential coupling
\begin{equation}
\delta W = \lambda S h^2 =\lambda S \epsilon^{ij}
h_i^{(1)}h_j^{(2)}~,
\label{lambda}
\end{equation}
where the chiral electroweak higgs superfield
$h=(h^{(1)}, h^{(2)})$ belongs to a
$(1,2,2)_{0}$ representation of $G_{LR}$ and
$\lambda$ can again be made positive. The scalar
potential which results from the
terms in Eqs.(\ref{W}) and (\ref{lambda}) is
(for canonical K\"ahler potential):
\begin{eqnarray}
V= | \kappa l^c\bar l^{c} + \lambda h^2 -
\kappa M^2|^2 +(m_{3/2}^2 + \kappa^2 |\bar l^c|^2
+ \kappa^2 |l^c|^2 + \lambda^2 |h|^2)|S|^2 +
m_{3/2}^2(|\bar l^c|^2
\nonumber\\
+ |l^c|^2 + |h|^2) +
\left (Am_{3/2}S( \kappa l^c\bar l^{c} +
\lambda h^2 - \kappa M^2) + 2\kappa m_{3/2} M^{2}S+
{\rm h. c.} \right )~,
\label{V}
\end{eqnarray}
where $m_{3/2}$ is the universal scalar mass (gravitino
mass) and $A$ the universal coefficient of the trilinear
soft terms. For exact supersymmetry ($m_{3/2}
\rightarrow 0$), the vacua are \cite{dls} at
\begin{equation}
S = 0,~~~\kappa l^c\bar l^{c} + \lambda h^2 =
\kappa M^2,~~~
l^c = e^{i\phi}\bar l^{c*}~~~h^{(1)}_i =
e^{i\theta}\epsilon_{ij}h^{(2)j*},
\end{equation}
where the last two conditions arise from the requirement
of D flatness. We see that there is a twofold degeneracy
of the vacuum which is lifted by supersymmetry breaking.
We get two degenerate (up to $m_{3/2}^{4}$) ground
states ($\kappa\neq\lambda$): the desirable (`good')
vacuum at $h = 0$ and $l^c\bar l^{c} = M^2$ and the
undesirable (`bad') one at $h \neq 0$ and
$l^c\bar l^{c} = 0$. They are separated by a potential
barrier of order $M^{2}m_{3/2}^{2}~$.
\par
To leading order in supersymmetry breaking, the term of
the potential $V$ in Eq.(\ref{V}) proportional to $A$
vanishes, but a destabilizing tadpole term for $S$
remains:
\begin{equation}
2\kappa m_{3/2}M^{2}S + {\rm h.c.}~.
\label{tadpole}
\end{equation}
This term together with the mass term of $S$
(evaluated at the `good' vacuum) give
$\langle S\rangle \approx - m_{3/2}/\kappa$ which,
substituted in Eq.(\ref{lambda}), generates \cite{dls}
a $\mu$ term with
\begin{equation}
\mu =\lambda \langle S\rangle \approx -
\frac{\lambda}{\kappa}m_{3/2}~.
\label{mu}
\end{equation}
Thus, coupling $S$ to the higgses can lead to the
resolution of the $\mu$ problem.
\par
The model can be extended \cite{dls} to include matter
fields too. The superpotential has the most general form
respecting the $G_{LR}$ gauge symmetry and a global
$U(1)$ R-symmetry. Baryon number is automatically
implied by this R-symmetry to all orders in the
superpotential, thereby guaranteeing the stability of
proton.
\subsection{The Inflationary Trajectory}
\label{subsec:inf}
\par
The model has \cite{lss,dls} a built-in inflationary
trajectory parametrized by $|S|$, $|S| > S_c=M$ for
$\lambda>\kappa$ (see below). All other fields vanish
on this trajectory. The $F_S$ term is constant providing
a constant tree level vacuum energy density
$\kappa^2 M^4$, which is responsible for
inflation. One-loop radiative corrections (from the mass
splitting in the supermultiplets $l^c$, $\bar l^{c}$
and $h$) generate a logarithmic slope \cite{dss} along
the inflationary trajectory which drives the inflaton
toward the minimum. For $|S| \leq S_c=M$, the $l^c$,
$\bar l^{c}$ components become tachyonic and the system
evolves towards the `good' supersymmetric minimum at
$h=0$, $l^c=\bar l^c=M$ (for $\kappa > \lambda$, $h$
is destabilized first and the system would have evolved
towards the `bad' minimum at $h \neq 0$, $l^c=
\bar l^c = 0$). For all values of the parameters
considered here, inflation continues at least till $|S|$
approaches the instability at $|S|=S_c$ as one deduces
from the slow roll conditions \cite{dss,lazarides}. The
cosmic microwave quadrupole anisotropy can be calculated
\cite {dss} by standard methods and turns out to be
\begin{equation}
\left(\frac{\delta T}{T}\right)_{Q} \approx
\frac{32 \pi^{5/2}}
{3\sqrt{5}}\left(\frac{M}{M_P}\right)^{3}
\kappa^{-1}x_{Q}^{-1}\Lambda (x_{Q})^{-1}~,
\label{anisotropy}
\end{equation}
\begin{eqnarray*}
\Lambda(x)=\left(\frac{\lambda}
{\kappa}\right)^{3}\left[\left(\frac{\lambda}
{\kappa}x^2-1\right)\ln \left(1-\frac{\kappa}
{\lambda}x^{-2}\right)
+\left(\frac{\lambda}{\kappa}x^2+1\right)
\ln \left(1+\frac{\kappa}
{\lambda}x^{-2}\right)\right]
\end{eqnarray*}
\begin{equation}
+(x^2-1)\ln (1-x^{-2})+(x^2+1)\ln (1+x^{-2})~,
\label{temp}
\end{equation}
with $x=|S|/S_c$ and $S_Q$ being the value of $|S|$
when the present horizon scale crossed outside the
inflationary horizon. (Notice that here we had to
replace the contribution to the effective potential
in Eq.(\ref{eq:veffexact}) from the $\phi$,
$\bar{\phi}$ supermultiplets of the model in
Sec.\ref{subsec:susy} by the contribution from the
$l^c$, $\bar l^{c}$ and $h$ supermultiplets.) The
number of e-foldings experienced by the universe between
the time the quadrupole scale exited the horizon and the
end of inflation is
\begin{equation}
N_Q \approx 32 \pi^3 \left(\frac{M}{M_P}\right)^2
\kappa^{-2}\int_{1}^{x_{Q}^{2}}\frac{dx^2}{x^2}
\Lambda(x)^{-1}~.
\label{efoldings}
\end{equation}
The spectral index of density perturbations turns out
to be very close to unity.
\subsection{`Reheating' and Leptogenesis}
\label{subsec:reheatlepto}
\par
After reaching the instability at $|S|=S_c$, the system
continues \cite{bl} inflating for another e-folding or
so reducing its energy density by a factor of about $2-3$~.
It then rapidly settles into a regular oscillatory phase
about the vacuum. Parametric resonance is safely ignored
in this case \cite{bl}. The inflaton (oscillating system)
consists of the two complex scalar fields $S$ and
$\theta=(\delta \phi + \delta\bar{\phi})/\sqrt{2}$,
where $\delta \phi = \phi - M$, $\delta \bar{\phi} =
\bar{\phi} - M$, with mass $m_{infl}=\sqrt{2}\kappa M$
($\phi$, $\bar{\phi}$ are the neutral components of
$l^c$, $\bar l^{c}$).
\par
The scalar fields $S$ and $\theta$ predominantly decay
into electroweak higgsinos and higgses respectively with
a common decay width $\Gamma_{h}=(1/16\pi)\lambda^{2}
m_{infl}$, as one can easily deduce from the couplings
in Eqs.(\ref{W}) and (\ref{lambda}). Note, however,
that $\theta$ can also decay to right handed neutrinos
$\nu^c$ through the nonrenormalizable superpotential
term
\begin{equation}
\frac{M_{\nu^c}}{2M^{2}}\bar{\phi}
\bar{\phi} \nu^c \nu^c~,
\label{majorana}
\end{equation}
allowed by the gauge and R- symmetries of the model
\cite{lss,dls}. Here, $M_{\nu^c}$ denotes the Majorana
mass of the relevant $\nu^c$. The scalar $\theta$ decays
preferably into the heaviest $\nu^c$ with $M_{\nu^{c}}
\leq m_{infl}/2$. The decay rate is given by
\begin{equation}
\Gamma_{\nu^c} \approx \frac{1}{16\pi}~\kappa^2
m_{infl}~ \alpha^2 (1-\alpha^2)^{1/2}~,
\label{decayneu}
\end{equation}
where $0\leq \alpha=2M_{\nu^c}/m_{infl} \leq 1$.
The subsequent decay of these $\nu^{c}$ 's produces
a primordial lepton number \cite{lepto} which is then
partially converted to the observed BAU through
electroweak sphaleron effects.
\par
The energy densities $\rho_{S}$, $\rho_{\theta}$,
and $\rho_{r}$ of the oscillating fields $S$, $\theta$,
and the `new' radiation produced by their decay to
higgsinos, higgses and $\nu^c$ 's are controlled by the
equations:
\begin{equation}
\dot{\rho}_{S}=-(3H+\Gamma_{h})\rho_{S}~,
~\rho_{\theta}(t)=\rho_{S}(t)
e^{-\Gamma_{\nu^c}(t-t_0)}~,
\label{infdensity}
\end{equation}
\begin{equation}
\dot{\rho}_{r}=-4H\rho_{r}+\Gamma_{h}\rho_{S}+
(\Gamma_{h}+\Gamma_{\nu^{c}})\rho_{\theta}~,
\label{raddensity}
\end{equation}
where
\begin{equation}
H=\frac{\sqrt{8\pi}}{\sqrt{3}M_P}~(\rho_{S}+
\rho_{\theta}+\rho_{r})^{1/2}
\label{hubble}
\end{equation}
is the Hubble parameter and overdots denote derivatives
with respect to cosmic time $t$. The cosmic time at the
onset of oscillations is taken $t_0\approx 0$. The
initial values of the various energy densities are taken
to be $\rho_{S}(t_0)=\rho_{\theta}(t_0)\approx
\kappa^{2}M^{4}/6$, $\rho_{r}(t_{0})=0$. The `reheat'
temperature $T_{r}$ is calculated from the equation
\begin{equation}
\rho_{S}+\rho_{\theta}=\rho_{r}=
\frac{\pi^2}{30}~g_{*}T_{r}^{4}~,
\label{reheat}
\end{equation}
where the effective number of massless degrees of
freedom is $g_{*}$=228.75 for MSSM.
\par
The lepton number density $n_{L}$ produced by the
$\nu^{c}$ 's satisfies the evolution equation:
\begin{equation}
\dot{n}_{L}=-3Hn_{L}+2\epsilon\Gamma_{\nu^c}
n_{\theta}~,
\label{lepton}
\end{equation}
where $\epsilon$ is the lepton number produced per
decaying right handed neutrino and the factor of 2 in
the second term of the rhs comes from the fact that we
get two $\nu^c$ 's for each decaying scalar $\theta$
particle. The `asymptotic' ($t\rightarrow 0$) lepton
asymmetry turns out to be
\begin{equation}
\frac{n_{L}(t)}{s(t)}\sim 3
\left(\frac{15}{8}\right)^{1/4}\pi^{-1/2}
g_{*}^{-1/4}m_{infl}^{-1}~
\frac{\epsilon\Gamma_{\nu^c}}
{\Gamma_{h}+\Gamma_{\nu^c}}~\rho_{r}^{-3/4}
\rho_{S}e^{\Gamma_{h}t}~.
\label{leptonasymmetry}
\end{equation}
\par
Assuming hierarchical light neutrino masses, we take
$m_{\nu_{\mu}}\approx 2.6\times 10^{-3}~\rm{eV}$
which is the central value of the $\mu$-neutrino mass
coming from the small angle MSW resolution of the solar
neutrino problem \cite{smirnov}. The $\tau$-neutrino
mass is taken $m_{\nu _{\tau}}\approx 7\times 10^{-2}
~\rm{eV}$, the central value from SuperKamiokande
\cite{superk}. Recent analysis \cite{giunti} of
the results of the CHOOZ experiment shows that
the oscillations of solar and atmospheric neutrinos
decouple. We thus concentrate on the two heaviest
families ignoring the first one. Under these
circumstances, the lepton number generated per
decaying $\nu^c$ is \cite{lazarides,neu}
\begin{equation}
\epsilon=\frac{1}{8\pi}~g
\left(\frac{M_3}{M_2}\right)
~\frac{{\rm c}^{2}{\rm s}^{2}\
\sin 2\delta \
(m_{3}^{D}\,^{2}-m_{2}^{D}\,^{2})^{2}}
{|\langle h^{(1)}\rangle|^{2}~(m_{3}^{D}\,^{2}\
{\rm s}^{2}\ +
\ m_{2}^{D}\,^{2}{\rm \ c^{2}})}~,
\label{epsilon}
\end{equation}
where $g(r)=r\ln (1 + r^{-2})$~,
$|\langle h^{(1)}\rangle|\approx 174~\rm{GeV}$,
${\rm c}=\cos \theta ,\ {\rm s}=\sin \theta $,
and $\theta$ ($0\leq \theta \leq \pi /2$) and
$\delta$ ($-\pi/2\leq \delta <\pi/2 $) are the
rotation angle and phase which diagonalize the
Majorana mass matrix of $\nu^{c}$ 's with eigenvalues
$M_2$, $M_3$ ($\geq 0$). The `Dirac' mass matrix
of the neutrinos is considered diagonal with
eigenvalues $m_{2}^{D}$, $m_{3}^{D}$ ($\geq 0$).
\par
For the range of parameters considered here, the
scalar $\theta$ decays into the second heaviest right
handed neutrino with mass $M_{2}$ ($<M_{3}$) and,
thus, $M_{\nu^{c}}$ in Eqs.(\ref{majorana}) and
(\ref{decayneu}) should be
identified with $M_{2}$. Moreover, $M_{3}$ turns out
to be bigger than $m_{infl}/2$ as it should. We will
denote the two positive eigenvalues of the light neutrino
mass matrix by $m_{2}$ (=$m_{\nu _{\mu }}$), $m_{3}$
(=$m_{\nu _{\tau }}$) with $m_{2}\leq m_{3}$. All the
quantities here (masses, rotation angles and phases) are
`asymptotic' (defined at the grand unification scale
$M_{GUT}$).
\par
The determinant and the trace invariance of
the light neutrino mass matrix imply\cite{neu}
two constraints on the (asymptotic) parameters which
take the form:
\begin{equation}
m_{2}m_{3}\ =\ \frac{\left( m_{2}^{D}m_{3}^{D}
\right) ^{2}}{M_{2}\ M_{3}}~,
\label{determinant}
\end{equation}
\begin{eqnarray*}
m_{2}\,^{2}+m_{3}\,^{2}\ =\frac{\left( m_{2}^{D}\,\,
^{2}{\rm c}^{2}+m_{3}^{D}\,^{2}{\rm s}^{2}\right)
^{2}}{M_{2}\,^{2}}+
\end{eqnarray*}
\begin{equation}
\ \frac{\left( m_{3}^{D}\,^{2}{\rm c}^{2}+m_{2}^{D}\,
^{2}{\rm s}^{2}\right)^{2}}{M_{3}\,^{2}}+\
\frac{2(m_{3}^{D}\,^{2}-m_{2}^{D}\,^{2})^{2}
{\rm c}^{2}{\rm s}^{2}\,{\cos 2\delta }}
{M_{2}\,M_{3}}~\cdot
\label{trace}
\end{equation}
\par
The $\mu-\tau$ mixing angle $\theta _{23}$
(=$\theta _{\mu\tau}$) lies \cite{neu} in the range
\begin{equation}
|\,\varphi -\theta ^{D}|\leq \theta _{23}\leq
\varphi +\theta ^{D},\ {\rm {for}\ \varphi +
\theta }^{D}\leq \ \pi /2~,~~~~~
\label{mixing}
\end{equation}
where $\varphi$ ($0\leq \varphi \leq \pi /2$) is the
rotation angle which diagonalizes the light neutrino mass
matrix, and $\theta ^{D}$ ($0\leq \theta ^{D} \leq
\pi /2$) is the `Dirac' (unphysical) mixing angle in
the $2-3$ leptonic sector defined in the absence of the
Majorana masses of the $\nu^{c}$ 's.
\par
Assuming approximate $SU(4)_{c}$ symmetry, we get the
asymptotic (at $M_{GUT}$) relations:
\begin{equation}
m_{2}^{D}\approx m_{c}\ ,
\ m_{3}^{D}\approx \ m_{t}\ ,
\ \sin\theta ^{D}\approx |V_{cb}|~.
\label{asympt}
\end{equation}
Renormalization effects, for MSSM spectrum and
$\tan \beta \approx m_{t}/m_{b}$, are incorporated
\cite{neu} by substituting in the above
formulas the values: $m_{2}^{D}\approx 0.23~{\rm GeV}$,
$\ m_{3}^{D}\approx 116$ GeV and $\sin \theta ^{D}
\approx 0.03$. Also, $\tan^{2} 2 \theta _{23}$
increases by about 40\% from $M_{GUT}$ to $M_{Z}$.
\par
We take a specific MSSM framework \cite{als}
where the three Yukawa couplings of the third generation
unify `asymptotically' and, thus,
$\tan \beta \approx m_{t}/m_{b}$. We choose the
universal scalar mass (gravitino mass) $m_{3/2}
\approx 290~{\rm{GeV}}$ and the universal gaugino mass
$M_{1/2} \approx 470~{\rm{GeV}}$. These values
correspond \cite{asw} to $m_{t}(m_{t})\approx 166~
{\rm{GeV}}$ and $m_{A}$ (the tree level CP-odd scalar
higgs mass) =$M_{Z}$. The ratio $\lambda/\kappa$ is
evaluated \cite{carena} from
\begin{equation}
\frac{\lambda}{\kappa}=\frac{|\mu|}{m_{3/2}}\approx
\frac{M_{1/2}}{m_{3/2}}\left(1-\frac{Y_{t}}
{Y_{f}}\right)^{-3/7}\approx 3.95~,
\label{renorm}
\end{equation}
where $Y_{t}=h_{t}^2\approx 0.91$ is the square of the
top-quark Yukawa coupling and $Y_{f}\approx 1.04$ is the
weak scale value of $Y_{t}$ corresponding to `infinite'
value at $M_{GUT}$.
\par
Eqs.(\ref{anisotropy})-(\ref{efoldings}) can now be
solved, for $(\delta T/T)_{Q} \approx 6.6\times 10^{-6}$
from COBE, $N_Q \approx 50$ and any value of $x_{Q}>1$.
Eliminating $x_{Q}$, we obtain $M$ as a function of
$\kappa$ depicted in Fig.\ref{reheating}. The evolution
Eqs.(\ref{infdensity})-(\ref{hubble}) are solved
for each value of $\kappa$. The parameter $\alpha^{2}$
in Eq.(\ref{decayneu}) is taken equal to 2/3. This
choice maximizes the decay width of the inflaton
to $\nu^{c}$ 's and, thus, the subsequently produced
lepton asymmetry. The `reheat' temperature is then
calculated from Eq.(\ref{reheat}) for each value of
$\kappa$. The result is again depicted in
Fig.\ref{reheating}.
\par
The mass of the second heaviest $\nu^c$, into which the
scalar $\theta$ decays partially, is given by $M_{2}=
M_{\nu^{c}}=\alpha m_{infl}/2$ and $M_{3}$ is found
from the `determinant' condition in Eq.(\ref{determinant}).
The `trace' condition in Eq.(\ref{trace}) is then solved
for $\delta(\theta )$ which is subsequently substituted
in Eq.(\ref{epsilon}) for $\epsilon$. The leptonic
asymmetry as a function of the angle $\theta$ can be found
from Eq.(\ref{leptonasymmetry}). For each value of
$\kappa$, there are two values of $\theta$ satisfying the
low deuterium abundance constraint $\Omega _{B}h^{2}
\approx 0.025$. (These values of $\theta$ turn out to be
quite insensitive to the exact value of $n_{B}/s$.) The
corresponding $\varphi$ 's are then found and the allowed
region of the mixing angle $\theta _{\mu\tau}$
in Eq.(\ref{mixing}) is determined for each $\kappa$.
Taking into account renomalization effects and
superimposing all the permitted regions, we obtain the
allowed range of $\sin^{2} 2 \theta _{\mu\tau}$
as a function of $\kappa$, shown in Fig.\ref{angle}.
We observe that $\sin^{2} 2 \theta _{\mu\tau}
\stackrel{_{>}}{_{\sim }} 0.8$ (from SuperKamiokande
\cite{superk}) corresponds to $1.2\times 10^{-6}
\stackrel{_{<}}{_{\sim }}\kappa
\stackrel{_{<}}{_{\sim }}3.4\times 10^{-6}$ which
is rather small. (Fortunately, supersymmetry protects it
from radiative corrections.)
\par
The corresponding values of $M$ and $T_{r}$ can be read
from Fig.\ref{reheating}. One finds that $1.4\times 10^{15}
~{\rm{GeV}}\stackrel{_{<}}{_{\sim }} M
\stackrel{_{<}}{_{\sim }}2\times 10^{15}~{\rm{GeV}}$
and $1.8\times 10^{7}~{\rm{GeV}}\stackrel{_{<}}{_{\sim }}
T_{r}\stackrel{_{<}}{_{\sim }}8.7\times 10^{7}
~{\rm{GeV}}$. We observe that $M$ turns out to be somewhat
smaller than the MSSM unification scale $M_{GUT}$. (It is
anticipated that $G_{LR}$ is embedded in a grand unified
theory.) The `reheat' temperature, however, satisfies the
gravitino constraint
($T_{r}\stackrel{_{<}}{_{\sim}} 10^{9}~{\rm{GeV}}$).
Note that, for the values of the parameters
chosen here, the lightest supersymmetric particle (LSP)
is \cite{asw} an almost pure bino with mass $m_{LSP}
\approx 0.43M_{1/2}\approx 200~{\rm{GeV}}$
and can, in principle, provide the cold dark matter of the
universe. On the contrary, there is no hot dark matter
candidate, in the simplest scheme.
\par
In conclusion, we have shown that, in a supersymmetric
model based on a left-right symmetric gauge group and
leading `naturally' to hybrid inflation, the $\mu$
problem can be easily solved . The observed BAU is
produced via a primordial leptogenesis. For masses of
$\nu_{\mu}$, $\nu_{\tau}$ from the small angle MSW
resolution of the solar neutrino puzzle and SuperKamiokande,
maximal $\nu_{\mu}-\nu_{\tau}$ mixing can be achieved.
The required values of the coupling constant $\kappa$ are,
however, quite small ($\sim 10^{-6}$).
\vspace{0.5cm}
This work is supported by E.U. under TMR contract
No. ERBFMRX--CT96--0090.
\def\ijmp#1#2#3{{ Int. Jour. Mod. Phys. }
{\bf #1~}(19#2)~#3}
\def\pl#1#2#3{{ Phys. Lett. }{\bf B#1~}(19#2)~#3}
\def\zp#1#2#3{{ Z. Phys. }{\bf C#1~}(19#2)~#3}
\def\prl#1#2#3{{ Phys. Rev. Lett. }{\bf #1~}(19#2)~#3}
\def\rmp#1#2#3{{ Rev. Mod. Phys. }{\bf #1~}(19#2)~#3}
\def\prep#1#2#3{{ Phys. Rep. }{\bf #1~}(19#2)~#3}
\def\pr#1#2#3{{ Phys. Rev. }{\bf D#1~}(19#2)~#3}
\def\np#1#2#3{{ Nucl. Phys. }{\bf B#1~}(19#2)~#3}
\def\mpl#1#2#3{{ Mod. Phys. Lett. }{\bf #1~}(19#2)~#3}
\def\arnps#1#2#3{{ Annu. Rev. Nucl. Part. Sci. }{\bf
#1~}(19#2)~#3}
\def\sjnp#1#2#3{{ Sov. J. Nucl. Phys. }{\bf #1~}
(19#2)~#3}
\def\jetp#1#2#3{{ JETP Lett. }{\bf #1~}(19#2)~#3}
\def\app#1#2#3{{ Acta Phys. Polon. }{\bf #1~}(19#2)~#3}
\def\rnc#1#2#3{{ Riv. Nuovo Cim. }{\bf #1~}(19#2)~#3}
\def\ap#1#2#3{{ Ann. Phys. }{\bf #1~}(19#2)~#3}
\def\ptp#1#2#3{{ Prog. Theor. Phys. }{\bf #1~}
(19#2)~#3}
\def\plb#1#2#3{{ Phys. Lett. }{\bf#1B~}(19#2)~#3}
\def\ibid#1#2#3{{ ibid. }{\bf#1~}(19#2)~#3}
\def\apjl#1#2#3{{ Astrophys. J. Lett. }{\bf#1~}(19#2)~#3}
|
\section{INTRODUCTION}
\label{sec_int}
In this paper, we study the scattering of surface plasmon polaritons
(SPP) by surface defects. SPP are $p$-polarized electromagnetic (EM)
waves bound to a dielectric-metal interface and caused by the surface
oscillations of the electron plasma of the metal.\cite{sp} They
propagate along the metal interface a distance of the order of the SPP
mean free path (ranging from microns in the visible to millimeters in
the infrared, of course depending also on the metal being considered),
undergoing scattering processes due to surface roughness. This
constitutes a classical problem of fundamental interest not only in the
case of individual defects (cf. Ref. \onlinecite{pmbg94} and references
therein), but also for periodically or randomly (or both)
distributed defects.\cite{pg96,prb96,kits96,prb97} Furthermore,
it is obviously crucial in any light scattering problem involving rough
metal surfaces where roughness-induced excitation of SPP occurs. This
has been explicitly shown in connection with either single defects
\cite{psg94,chema95,alb} or random corrugation,\cite{psg95,wo,mmm95,owm98}
the latter configuration being relevant to
the phenomenon of (SPP mediated) enhanced backscattering of light.
In addition to that, light-SPP coupling plays a central role in other
phenomena such as anomalous transmission through metal slabs with hole
arrays,\cite{ebbe98,schro98} surface-enhanced Raman scattering,
\cite{sers,vp96,jcp98} or biosensing.\cite{bio}
In recent years, the advent of near-field optical microscopy \cite{nf}
has opened up the possibility to study experimentally SPP in a direct
manner. Among the various configurations developed, photon scanning
tunneling microscopy \cite{pstm} (PSTM), basically exploiting SPP
excitation in the attenuated total reflection arrangement, has made it
possible to probe the SPP structure,\cite{marti,adam} localized SPP on
randomly rough surfaces,\cite{bozh95} and SPP resonances in fractal
colloid clusters \cite{tsai94} and single particles.\cite{nie97,klar98}
Moreover, PSTM images have been obtained by surface-enhanced Raman
scattering probing single molecules adsorbed on single nanoparticles.
\cite{nie97} PSTM in combination with direct-write lithography has
made it possible to create sub-micron defects on metal
surfaces.\cite{smol95}
Particularly relevant to the present work are the recent experimental
studies on SPP scattering by surface defects.\cite{smol,bozh,smol99}
These studies have shown evidence of drastically distinct scattering
properties depending on the defect size. Specifically, surface defects
favoring SPP reflection and light coupling, called SPP mirrors and
flashlights,\cite{smol} respectively, have been described, as well as
SPP microlenses and microcavities;\cite{bozh} SPP Bloch waves have also
been imaged in periodic arrays of surface defects.\cite{smol99}
Interestingly, the possibilities of artificially creating micro-optical
components for SPP have also been noted in these studies. Much has to be
done, however, from the theoretical standpoint. Quite recently, calculations
for circularly symmetric defects have successfully accounted for the
peculiar azimuthal dependence of the radiated pattern;\cite{snm97} in
addition, such calculations have been used to retrieve the surface profile.
\cite{pasc98} In the case of one-dimensional surface defects, preliminary
calculations have focused on the optimization of the defect size to obtain
SPP mirrors and so-called light-emitters.\cite{apl98} In this regard, it
is our purpose to address in detail the SPP scattering by one-dimensional
surface defects, including near-field and far-field calculations (along
with energy balance) and their dependence on defect size parameters. Thus
we expect not only to shed light on the experimental works mentioned above,
but also to find and predict related effects.
The physical system we consider here is a planar one-dimensional
metal surface with a one-dimensional defect. The surface corrugation
is modeled by using a local impedance boundary condition (IBC) on a flat
surface. The connection between surface impedance and real surface
corrugation has been recently demonstrated,\cite{aamibc} and its
validity to give accurate quantitative results has been shown
in numerical calculations of grating-induced SPP-photon coupling.
\cite{prb96} A scattering-theoretic formulation of the interaction of
an SPP with the surface roughness is developed by imposing the IBC
on the amplitude of the magnetic field in the vacuum region in the
form of a Rayleigh expansion. Upon solving the resulting integral
equation for the scattering amplitude, the magnetic field at any
point in the vacuum half-space can be calculated. We will focus on
the far field angular distribution and the surface field amplitudes
to determine, respectively, the total radiated energy $S$, and the
SPP reflection $R_{SP}$ and transmission $T_{SP}$ coefficients. By
numerical simulation calculations, these quantities are computed.
The paper is organized as follows. The theoretical formulation
is derived in Sec.~\ref{sec_the}, and some details pertaining to the
numerical procedure are given in the Appendix. In Sec.~\ref{sec_res}, we
show the results obtained for a single Gaussian defect and the influence
of defect width and height. Finally, Section~\ref{sec_con} summarizes the
conclusions drawn from this research.
\section{THEORY}
\label{sec_the}
\subsection{Scattering Equations}
\label{sec_the_se}
We study the scattering of a p-polarized SPP of frequency
$\omega$ propagating along a flat vacuum-metal interface ($x_3=0$)
by a one-dimensional obstacle (constant along the $x_2$-axis, see
Fig.~\ref{fig_plasmon}). Under these circumstances, the three-dimensional
electromagnetic problem can be cast into a two-dimensional scalar
problem in such a way that the single, nonzero component of the
magnetic field amplitude $H_2(x_1,x_3)$ is the solution of the
corresponding two-dimensional Helmholtz equation in the upper
(vacuum) and lower (metal) half spaces. The magnetic field in vacuum
is assumed to be the sum of an incoming SPP and a scattered field as
follows
\begin{eqnarray}
\lefteqn{H_2^>(x_1,x_3) =\exp[ik(\omega)x_1-\beta_o(\omega)x_3]} &&
\nonumber \\ && \phantom{+++}+\int_{-\infty}^{\infty}\frac{dq}{2\pi}
\, R(q,\omega)\exp[iqx_1+i\alpha_o(q,\omega)x_3] ,
\label{eq_h}
\end{eqnarray}
where
\begin{eqnarray}
k(\omega) & \equiv k^R(\omega)+ik^I(\omega)={\displaystyle
\frac{\omega}{c}\left(1-\frac{1}{\epsilon(\omega)}
\right)^{1/2},}
\label{eq_kw}
\end{eqnarray}
\begin{eqnarray}
\beta_0(\omega) & = & \left(k(\omega)^2-\frac{\omega^2}{c^2}
\right)^{1/2}=\frac{\omega}{c}[-\epsilon(\omega
)]^{-1/2},
\label{eq_bw}
\end{eqnarray}
and
\begin{mathletters}\begin{eqnarray}
\alpha_o(q,\omega) & = & \left(\frac{\omega^2}{c^2}-q^2
\right)^{1/2}\;\;\;\; |q|\leq\frac{\omega}{c} \\
& = & i\left(q^2-\frac{\omega^2}{c^2}\right)^{1/2}
\;\;\; |q|>\frac{\omega}{c}.
\end{eqnarray}\end{mathletters}
Note that the expressions for the SPP wavevector components $k(\omega)$
and $\beta_0(\omega)$ in vacuum apply in the limit $|\epsilon(\omega)|\gg
1$. This stems from the fact that the continuity conditions
across the interface are mapped onto a local IBC on the planar
surface $x_3=0$ in the form
\begin{eqnarray}
\lefteqn{\left.\frac{\partial}{\partial x_3} H_2^>(x_1,x_3)\right|_{x_3=0}}
&& \nonumber \\ &&
=-\frac{\omega}{c}\,\frac{1+s(x_1)}{[-\epsilon (\omega)]^{ 1/2}}
\;H_2^>(x_1,x_3)|_{x_3=0} ,
\label{eq_ibc}
\end{eqnarray}
where the superscript $>$ indicates the vacuum region,
$-(\omega/c)[-\epsilon (\omega )]^{-1/2}s(x_1)$ is the contribution to the
surface impedance associated with the obstacle, and $\epsilon(\omega)$
is the isotropic, frequency-dependent dielectric function of the metal.
The IBC has been widely used in the past to model the vacuum-metal
interface qualitatively, especially in the infrared region of the optical
spectrum. Furthermore, it has been recently proven to be quantitatively
accurate in calculations of grating-induced photon-SPP coupling \cite{prb96}
by using the connection between surface impedance and real corrugation
demonstrated in Ref. \onlinecite{aamibc}.
In order to calculate the scattering amplitude $R(q,\omega)$, we
substitute Eq. (\ref{eq_h}) into Eq. (\ref{eq_ibc}), and obtain the
integral equation $R(q,\omega )$ satisfies,
\begin{eqnarray}
R(q,\omega)=&& G_0(q,\omega)V\bbox(q|k(\omega)\bbox) \nonumber\\ &&
+G_0(q,\omega)\int_{-\infty}^{\infty}\! \frac{dp}{2\pi}
V(q|p)R(p,\omega),
\label{eq_ierq}
\end{eqnarray}
where
\begin{eqnarray}
G_0(q,\omega) & \equiv &
\frac{i\epsilon(\omega)}{\epsilon(\omega)\alpha_0(q,\omega)
+i(\omega/c)[-\epsilon(\omega)]^{1/2}}
\label{eq_gq}
\end{eqnarray}
is the Green's function of the SPP on the unperturbed surface
[$s(x_1)=0$]. We have also introduced the scattering potential
\begin{eqnarray}
V(q|p) & \equiv & \beta_0(\omega)\hat{s}(q-p),
\label{eq_vq}
\end{eqnarray}
with
\begin{eqnarray}
\hat{s}(Q)&=&\int_{-\infty}^{\infty}\! dx_1\; e^{-iQx_1}s(x_1),
\label{eq_sq}
\end{eqnarray}
to simplify the notation. Equation (\ref{eq_ierq}) can be rewritten in a
more convenient manner by substituting
\begin{eqnarray}
R(q,\omega) &=& G_0(q,\omega)T(q,\omega),
\label{eq_rtq}
\end{eqnarray}
into it, so that
\begin{eqnarray}
T(q,\omega)=&& V\bbox(q|k(\omega)\bbox) \nonumber \\ &&
+\int_{-\infty}^{\infty}\!
\frac{dp}{2\pi}V(q|p)G_0(p,\omega)T(p,\omega).
\label{eq_ietq}
\end{eqnarray}
Equation (\ref{eq_ietq}), along with Eqs. (\ref{eq_h}) and
(\ref{eq_rtq}), is the basis of our theoretical formulation.
In solving
Eq. (\ref{eq_ietq}), it is very important how we deal with the poles
appearing in the Green's function (\ref{eq_gq}). First, we rewrite the
latter in the form
\begin{eqnarray}
G_0(q,\omega) & = & C(q,\omega)\left(
\frac{1}{q-k(\omega)}-\frac{1}{q+k(\omega)}\right),
\end{eqnarray}
with
\begin{eqnarray}
C(q,\omega) & \equiv & \frac{\epsilon(\omega)\alpha_0(q,\omega)-
i(\omega/c)[-\epsilon(\omega)]^{1/2}}{2i\epsilon(\omega) k(\omega)}.
\end{eqnarray}
We now assume that the metal dielectric function is given by
Drude's expression
\begin{eqnarray}
\epsilon(\omega) &=& 1-\frac{\omega_p^2}{\omega^2} ,
\end{eqnarray}
where $\omega_p$ is the plasma frequency, in the absence of
absorption losses.
Therefore, in light of Eq. (\ref{eq_kw}), we have to take the limit
$k^I(\omega)\rightarrow 0$ in Eq. (\ref{eq_gq}), to obtain
\begin{eqnarray}
G_0(q,\omega) &&= C(q,\omega)\left(
\left.\frac{1}{q-k^R(\omega)}\right|_P-\left.
\frac{1}{q+k^R(\omega)}\right|_P\right. \nonumber \\ &&
\left.\phantom{\frac{1}{2}}+\pi i[\delta\bbox(q-k^R(\omega)\bbox)
+\delta\bbox(q+k^R(\omega)\bbox)]\right) .
\label{eq_gqri}
\end{eqnarray}
The first two terms on the right-hand side of Eq. (\ref{eq_gqri}) have
meanings in the Cauchy's principal value sense, whereas the last
two terms are delta functions. Once Eq. (\ref{eq_ietq}) is solved for
$T(q,\omega)$ [we will see below how to do so numerically with
the help of Eq. (\ref{eq_gqri})], we proceed to calculate the
electric and magnetic near fields, the SPP-photon coupling, and the
SPP reflection and transmission coefficients in the following manner.
\subsection{Near Field}
\label{sec_the_nf}
The magnetic field at any point in the vacuum half-space can be
straightforwardly calculated from Eq. (\ref{eq_h}), upon recalling
Eq. (\ref{eq_rtq}), which relates $T(q,\omega)$ with the scattering
amplitude $R(q,\omega)$. Then the electric field components in vacuum are
easily written also as functions of $R(q,\omega)$ by means of a Maxwell
curl equation as follows:
\begin{mathletters}\label{eq_e}\begin{eqnarray}
E_1^>(x_1,x_3) = &&\frac{c}{\omega}i\beta_0(\omega)
\exp[ik(\omega)x_1-\beta_0(\omega)x_3] \nonumber \\ &&
+ \frac{c}{\omega}\int_{-\infty}^{\infty}\frac{dq}{2\pi}
\,\alpha_0(q,\omega) R(q,\omega) \nonumber \\ && \times
\exp[iqx_1+i\alpha_0(q,\omega)x_3] \\
E_2^>(x_1,x_3) = && 0 \\
E_3^>(x_1,x_3) = &&-\frac{c}{\omega}k(\omega)
\exp[ik(\omega)x_1-\beta_0(\omega)x_3] \nonumber \\ &&
- \frac{c}{\omega}\int_{-\infty}^{\infty}
\frac{dq}{2\pi}\, qR(q,\omega) \nonumber \\ && \times
\exp[iqx_1+i\alpha_0(q,\omega)x_3].
\end{eqnarray}\end{mathletters}
The time-averaged Poynting vector thus reads:
\begin{eqnarray}
<{\bf S}> = \frac{c}{8\pi} \Re \left({\bf E}\times{\bf H}^*\right) =
\frac{c}{8\pi} \Re \left(-E_3 H_2^*,0,E_1 H_2^*\right) ,
\label{eq_pv}
\end{eqnarray}
where $\Re$ denotes the real part and the asterisk the complex
conjugate.
\subsection{Radiated energy}
\label{sec_the_s}
The total power carried away from the surface in the form of
volume electromagnetic waves propagating in the vacuum region
above it, per unit length of the system along the $x_2$-axis is
\begin{eqnarray}
P_{sc}= & & \int_{-\infty}^{\infty}\! dx_1\; <S^{(sc)}_3> \nonumber \\
= &&\frac{c^2}{8\pi\omega}
\int_{-\omega/c}^{\omega/c}
\frac{dq}{2\pi}\alpha_0(q,\omega)\mid\!R(q,\omega)\!\mid^2 .
\label{eq_psc}
\end{eqnarray}
Note that only the scattered field contribution to the $x_3$-component of
the time-averaged Poynting vector is used. Equation (\ref{eq_psc}) must
be normalized by the power carried by the incident SPP
per unit length along the $x_2$-axis
\begin{eqnarray}
P_{inc}= & & \int_0^{\infty}\! dx_3\; <S^{(inc)}_1>
=\frac{c^2k(\omega)}{16\pi\omega\beta_0(\omega)} ,
\label{eq_pin}
\end{eqnarray}
where $<S^{(inc)}_1>$ is the $x_1$-component of the time-averaged Poynting
vector of the incident SPP. Then the total, normalized scattered power
$S$ is given by
\begin{eqnarray}
S &= {\displaystyle\frac{P_{sc}}{P_{inc}}} &
=\int_{-\pi/2}^{\pi/2}
\! d\theta_s \frac{\partial R}{\partial\theta_s},
\label{eq_s}
\end{eqnarray}
where
\begin{eqnarray}
\frac{\partial R}{\partial\theta_s} =& &{\displaystyle\frac{1}{2
\pi}\frac{\beta_0(\omega)}{2k(\omega)}\;\alpha_0^2\left(
q=\frac{\omega}{c}\sin\theta_s\right)} \nonumber\\ && \times\left|
R\left(q=\frac{\omega}{c}\sin\theta_s\right)\right|^2 ,
\label{eq_drc}
\end{eqnarray}
is the differential reflection coefficient (DRC), namely, the fraction
of the energy of the incident SPP that is scattered into an angular
region of width $d\theta_s$ about the scattering direction $\theta_s$
where the scattering angle $\theta_s$ is measured clockwise with respect
to the $x_3$-axis (see Fig.~\ref{fig_plasmon}).
\subsection{Reflection and transmission coefficients}
\label{sec_the_rt}
In order to evaluate the amplitude of the reflected and transmitted SPP,
we study the behavior of $H_2^>(x_1,x_3)$, Eq. (\ref{eq_h}), with the help
of Eqs. (10) and (15), on the surface $x_3=0$. At this point, great care
has to be taken when calculating the contribution to the scattered field
in Eq. (2) from the Cauchy principal value integrals arising from the first
two terms on the right-hand side of Eq. (15). We assume that the
obstacle has a finite extent and is centered about $x_1=0$. If
we focus on the regions $x_1\ll 0$ and $x_1\gg 0$ far from the
obstacle, it can be shown, by working out the contributions from
those integrals in the complex $q$-space with the help of Cauchy's
theorem,\cite{cauchy} that the magnetic field is given by:
\begin{mathletters}\label{eq_hxinf}\begin{eqnarray}
H_2^>(x_1,x_3=0) =& &\exp[ik^R(\omega)x_1] \nonumber \\ & &
+r(\omega)\exp[-ik^R(\omega)x_1], x_1\ll 0 \\
= && t(\omega)\exp[ik^R(\omega)x_1], \phantom{+-} x_1\gg 0,
\end{eqnarray}\end{mathletters}
where the amplitudes of the reflected and transmitted SPP,
$r(\omega)$ and $t(\omega)$, respectively, are
\begin{mathletters}\label{eq_spprt}\begin{eqnarray}
r(\omega) &=& iT\bbox(-k^R(\omega),\omega\bbox)\; C\bbox(-k^R(\omega),
\omega\bbox) \\
t(\omega) &=& 1+iT\bbox(k^R(\omega),\omega\bbox)\; C\bbox(k^R(\omega),
\omega\bbox) .
\end{eqnarray}\end{mathletters}
Equations (\ref{eq_hxinf}) manifest the fact that, away from the
obstacle, only the incident and reflected SPP (on the left-hand side,
see Fig. 1) and the transmitted SPP (on the right-hand side)
propagate along the interface. The corresponding reflection
and transmission coefficients are
\begin{mathletters}\label{eq_spprt2}\begin{eqnarray}
R(\omega) &=& \mid\! r(\omega)\!\mid^2 \\
T(\omega) &=& \mid\! t(\omega)\!\mid^2 .
\end{eqnarray}\end{mathletters}
\subsection{Numerical Calculations}
\label{sec_the_nc}
The integral equation (\ref{eq_ietq}) is numerically solved by
converting it into a matrix equation through a quadrature scheme.
The details are given in the Appendix. It should be pointed out
that the discretization $q$-mesh is chosen in such a way that
$q=\pm k^R(\omega)$ are always points on the mesh, as required
by Eq. (\ref{eq_spprt}). In addition, the discretization is not
regular: the density of $q$-points around the poles at
$q=\pm k^R(\omega)$ is considerably larger ($\Delta q\approx
10^{-4}\omega/c$) than it is either in the radiative region
$|q|\leq\omega/c$ or in the non-radiative region away from the
poles ($\Delta q\approx 10^{-2}\omega/c$). The number $N$ of
$q$-points needed in the numerical procedure depends not only on
the accuracy required to sample the pole regions, but also on
the explicit form of the obstacle, which enters in the calculation
through its Fourier-transform in Eq. (\ref{eq_sq}). Throughout this
work, typically $N=2600$, except for the larger defects for which up to
$N=4000$ points are employed. The convergence of the numerical results
with increasing $N$ has been checked in the most unfavorable cases.
\section{RESULTS AND DISCUSSION}
\label{sec_res}
Note that up to now no restrictions have been imposed on the shape
of the obstacle apart from its having a finite extent along the
$x_1$-axis. Its surface impedance function $s(x_1)$ is connected to the
actual surface profile defined by $x_3 = f(x_1)$ through\cite{aamibc}
\begin{eqnarray}
s(x) & = & -\frac{1-\epsilon(\omega)}{d(\omega)\epsilon(\omega)}
[1-d^2(\omega)D^2]^{1/2} f(x_1) + O(f^2),
\label{eq_sh}
\end{eqnarray}
where $d(\omega)=(c/\omega)[-\epsilon(\omega)]^{-1/2}$ is the optical skin
depth, and $D\equiv d/dx_1$. In the case of small skin depths and surface
slopes $(dD)^2\ll 1$, the square root term on the rhs of Eq.
(\ref{eq_sh}) can be expanded as
\begin{eqnarray}
\lefteqn{[1-(dD)^2]^{1/2}= 1-\frac{1}{2} (dD)^2-\ldots} \nonumber \\
&& \phantom{++}-\frac{1.1.3\ldots(2n-3)}{2.4.6\ldots 2n}(dD)^{2n}
+O\bbox((dD)^{2n+2}\bbox).
\label{eq_ddn}
\end{eqnarray}
Then the Fourier transform of the surface impedance function, which is
needed in the calculation [cf. Eq. (\ref{eq_vq})], is related to the
Fourier transform $\hat{f}(Q)$ of the surface profile function through
\begin{eqnarray}
\lefteqn{\hat{s}(Q)= -\frac{1-\epsilon(\omega)}{d(\omega)
\epsilon(\omega)}\left(1-\frac{1}{2} [-i d(\omega)Q]^2\right.}
&& \nonumber \\ && \left.\phantom{++} -\frac{1}{8}[-i d(\omega)Q]^4
+O\bbox([-i d(\omega)Q]^6\bbox)\right)\hat{f}(Q).
\label{eq_sqf}
\end{eqnarray}
In what follows, we will restrict the analysis to a Gaussian
defect of $1/e$ half-width $a$ and height $h$:
\begin{eqnarray}
f(x_1) & = & h\exp\left(-x_1^2/a^2\right).
\end{eqnarray}
In addition, unless otherwise stated, we retain in Eq. (\ref{eq_ddn}),
and thus in Eq. (\ref{eq_sqf}), only the zeroth order term in the
expansion in powers of $[-i d(\omega)Q]^2$, as implicitly done in
Ref. \onlinecite{apl98}. Therefore the function $\hat{s}(Q)$ we will use
in our calculations is
\begin{mathletters}
\begin{eqnarray}
\hat{s}(Q) &=& \pi^{1/2} s_0
a\exp[-(aQ)^2/4],
\end{eqnarray}
with
\begin{eqnarray}
s_0 &=& -\frac{1-\epsilon(\omega)}{\epsilon(\omega)}
\frac{h}{d(\omega)}.
\end{eqnarray}
\label{eq_sqfg}
\end{mathletters}
It should be emphasized that the approximation involved in retaining only
the lowest order term in the expansion (\ref{eq_ddn}) affects only the
expression connecting the surface impedance with the real surface
profile, the scattering formulation being rigorous and energy conserving
(recall that losses are not accounted for) whatever the surface impedance
is. Nonetheless, inasmuch as we wish to be able to quantitatively relate
our results with real defect sizes, the effect of neglecting the higher
order terms in Eq. (\ref{eq_ddn}) has to be determined. We have thus
verified in the most unfavorable cases that including the first order
term in $[-id(\omega )Q]^2$ in Eq. (\ref{eq_sqf}) barely modifies our
calculations.
In order to establish the accuracy and efficiency of the numerical
calculations based on the formulation above, we first calculate the
function $T(q,\omega)$ [cf. Eq. (\ref{eq_ietq})], following the numerical
procedure outlined in Sec. \ref{sec_the} and the Appendix, for two
Gaussian defects of half-width $a/\lambda=0.1$ and heights
$h/\lambda=\pm 0.05$ (protuberance and indentation of equal height/depth),
where $\lambda$ is the wavelength of the SPP. From these results, the SPP
reflection and transmission coefficients are straightforwardly calculated
[cf. Eqs. (\ref{eq_spprt}) and (\ref{eq_spprt2})], along with the DRC
[cf. Eqs. (\ref{eq_rtq}) and (\ref{eq_drc})]. Furthermore, the magnetic
and electric fields at any point in the vacuum half-space can be
calculated from Eqs. (\ref{eq_h}) and (\ref{eq_e}) by using Eq.
(\ref{eq_rtq}). In Fig.~\ref{fig_hi} we present the results thus obtained
for the magnetic field intensities at the vacuum-metal interface in the
vicinity of the Gaussian defects, and for the angular distribution of the
scattered field in the far field. From the surface magnetic field in
Fig.~\ref{fig_hi}(a), it is evident that both surface defects reflect
back part of the incoming SPP, which interferes with the incoming SPP
giving rise to the oscillatory pattern to the left of the defect
(negative $x_1$-axis). Near the defect the magnetic field is perturbed.
The outgoing transmitted SPP is seen to the right of the defect. The SPP
reflection and transmission coefficients are: $R_{SP}=0.0025$ and
$T_{SP}=0.9825$ for the protuberance, and $R_{SP}=0.0041$ and
$T_{SP}=0.9728$ for the indentation.
In Fig.~\ref{fig_hi}(b) a fairly uniform angular distribution of the DRC
is observed (this will be discussed below). The total scattered power
calculated
from Eq. (\ref{eq_s}) is $S=0.0149$ for the protuberance, and $S=0.0231$
for the indentation. Energy conservation is thus satisfied
within a $0.01\%$ error.
We find that away from the vicinity of the defect, the magnetic field is
fully described by either the interference between incoming and reflected
SPP on the left-hand side, or by merely the transmitted SPP on the
right-hand side. This corroborates, as expected, our argument in
Sec.~\ref{sec_the_rt} leading to Eqs. (\ref{eq_hxinf}). The reflection and
transmission coefficients are thus calculated from Eqs. (\ref{eq_spprt})
and (\ref{eq_spprt2}).
\subsection{Energy balance dependence on defect size}
\label{sec_res_rts}
The question now arises naturally as to how efficient the surface defect
is in coupling the incoming SPP into the different outgoing channels
(either SPP or photons), or conversely, what the appropriate defect
parameters are that maximize or minimize those channels; this is crucial
for both an understanding of the scattering process and the design of
practical devices. To that end, we have studied the dependence of the
scattering coefficients $R_{SP}$, $T_{SP}$, and $S$ on the defect
half-width $a$ for both Gaussian protuberances and indentations of
different heights $h/\lambda=0.05$ and 0.2. The results are shown in
Fig.~\ref{fig_rts}. Several general features are evident from these
results.
First, SPP reflection is relevant only for very narrow defects,
$a<\lambda/5$, for either protuberances or indentations. Indeed there
is an optimum defect width for which $R_{SP}$ is maximum.\cite{apl98}
These defects are called {\it plasmon mirrors}.\cite{smol,apl98} For
increasing defect widths, protuberances and indentations begin to
behave differently, except for their negligible contribution to SPP
reflection. On the one hand, SPP transmission through protuberances
monotonically diminishes at the expense of radiation. The conversion
is steeper the higher the defect is.
Indentations, however, exhibit an oscillatory pattern with increasing
defect width, in such a way that radiation (SPP transmission)
increases (decreases), passes through a maximum (minimum), and then
tends asymptotically to 0 (1). The oscillation period, the defect width
that yields maximum radiation, and the value of this maximum, all
depend on the surface height. Note that both protuberances and
indentation may behave as {\it light-emitters}\cite{apl98} (high
SPP-light conversion efficiency) for an appropriate (and distinct)
range of defect parameters. Below we analyze in detail the behavior of
SPP mirrors and light-emitters.
\subsection{Narrow defects: SPP mirrors}
\label{sec_res_mir}
Surface defects playing the role of SPP mirrors have been studied
experimentally in PSTM configurations.\cite{smol} This phenomenon has been
analyzed for different defect shapes in Ref. \onlinecite{apl98}, where in
addition a simple analytical prediction is given through a
perturbation-theoretic argument. In the case of Gaussian-shaped defects,
the predicted half-width that yields maximum reflection is:
$a_{mirr}\approx [2^{1/2}k^R(\omega)]^{-1}$. Our numerical
results further corroborate this prediction, since it is seen in
Fig.~\ref{fig_rts} that $a_{mirr}/\lambda\approx 0.1$ no matter what the
defect height is [as long as this height does not exceed the range of
validity of Eq. (\ref{eq_sh})]. Nonetheless, the maximum SPP reflection
increases with the defect height, and is slightly larger for indentations.
The electric and magnetic near-field intensities for Gaussian defects,
placed at the origin, of width $a/\lambda=0.1$ and heights $h/\lambda=\pm
0.05$ (protuberance and indentation, respectively) are presented in
Figs.~\ref{fig_nfpm_p} and~\ref{fig_nfpm_i}, respectively,. The near-field
maps are quite similar in both cases. The oscillations to the left of the
obstacles clearly reveal the interference between the incident and
backscattered SPP, their period being $T\approx 2\pi/(2k^R)$, as expected,
and their contrast being related to $R_{SP}$. A bright region is seen to
the right of the defects that is due to the strong SPP transmission.
Poynting vector maps superimposed on the electric near-field intensity
maps confirm the description of the energy flow given above.
The corresponding angular distributions of scattered light (DRC) have been
shown in Fig.~\ref{fig_hi}(b). There are no significant qualitative
differences between protuberances and indentations, both yielding a fairly
structure-less angular dependence; quantitatively, an indentation leads to
stronger light coupling. The qualitative behavior is somewhat expected:
the same perturbation-theoretic argument predicting maximum SPP reflection
for a defect width that maximizes the scattering potential [cf. Eq.
(\ref{eq_vq})] at backscattering,\cite{apl98} leads to a mostly uniform
SPP coupling into EM waves in the radiative region ($|q|\leq\omega/c$).
\subsection{Wide protuberances: Total light-emitters}
\label{sec_res_wp}
The ability of Gaussian-shaped protuberances to couple SPP into light
has been pointed out in Ref. \onlinecite{apl98}. Here we analyze in detail
the conditions for protuberances and indentations alike to behave as
light-emitters with coupling efficiencies beyond 90$\%$, larger than that
reported in Ref. \onlinecite{apl98}. Figure~\ref{fig_rts} above illustrates
the discussion.
To begin with, let us focus on protuberances. For widths beyond those
producing significant SPP reflection (SPP mirrors), SPP-light conversion
increases monotonically whereas, as expected from energy conservation,
SPP transmission decays. This variation is faster for higher
protuberances. Indeed, the curves in Fig. \ref{fig_rts} (bottom) indicate
that $\lim_{a\rightarrow\infty}S=1$ even for the small protuberance.
We have found coupling efficiencies beyond 90$\%$ in the
case of $h/\lambda=0.2$ and $a/\lambda\geq 3.6$. In
Fig.~\ref{fig_nfle_p}, the electric and magnetic near-field
intensity maps for one such defect are shown. The absence of
oscillations to the left of the defect reveals that SPP reflection
is small; SPP transmission is small too (though considerably larger than
$R_{SP}$), as seen on the right-hand side of the defect.
A light beam is observed leaving the surface from the defect at
near-grazing scattering angles. This energy flow picture is further
corroborated by the angular distribution of the DRC shown in
Fig.~\ref{fig_ile}, with a maximum at $\theta_s\approx 74^{\circ}$.
Qualitatively, the fact that the metal protuberance enters the vacuum
half-space seems to favor the SPP-photon coupling (playing the role of a
{\it launching platform}).
\subsection{Wide indentations: Light-emitters and SPP total transmission}
\label{sec_res_wi}
In the case of wide indentations, however, the behavior of the
different outgoing channels differs from that for protuberances, and
exhibits a richer phenomenology. Upon increasing the width of the
indentation beyond the range of significant SPP reflection
(see Fig.~\ref{fig_rts}), SPP
transmission reaches a minimum value leading to maximum radiation, and
then slowly grows towards {\it total transmission} (no radiation) in
an oscillatory manner. The defect width that yields maximum radiation,
its value, and the oscillations depend on the defect height.
To understand such behavior, we plot in Figs.~\ref{fig_nfle_i_3}
and~\ref{fig_nfle_i4} the electric near-field intensity maps for the
higher defects ($h/\lambda=0.2$) of widths $a/\lambda=0.3$ and
$a/\lambda=4$, respectively. These widths correspond to the first
absolute and sixth subsidiary, maxima in Fig.~\ref{fig_rts} (bottom),
respectively. Both indentations give rise to
a negligible amount of SPP reflection (no oscillations to the left of
the defect), as expected from Fig.~\ref{fig_rts} (top). SPP transmission
(to the right of the defect) is very small for $a/\lambda=0.3$, but
a strong light beam at grazing scattering is observed ($S=94\%$). For
$a/\lambda=4$, although most of the energy goes into $T_{SP}=86.1\%$,
a small amount of radiation also at grazing scattering angles is seen
[recall that even though for this width a local maximum occurs in $S$,
its value is very small $S=13.3\%$, see Fig~\ref{fig_rts} (bottom)].
But what is very illustrative to the discussion on the behavior of the
outgoing channels is the near-field within the indentation (strictly
speaking, right on top of the indentation region, since we are using an
IBC on a flat surface). Oscillations are found therein, the number of
minima (one in Fig.~\ref{fig_nfle_i_3} and six in
Fig.~\ref{fig_nfle_i4}) being directly related to the position of
the corresponding maxima in the $S$ vs. $a$ curve [see
Fig.~\ref{fig_rts}]. Therefore, it can be inferred that the
oscillatory behavior of SPP transmission and conversion into light in
indentations is governed by a cavity-like effect.
In fact, the near field map (not shown here) in the
vicinity of any minimum-radiation indentation provides further evidence
for this suggestion.
As a consequence, the range of defect widths for which high coupling
efficiencies are encountered is far more restrictive for
indentations. In fact, in contrast with protuberances, only sufficiently
deep indentations ($h/\lambda\geq 0.2$) are capable of
producing radiation efficiencies $S>90\%$, and only for a narrow range
of parameters. It seems as if the indentation geometry would somehow
hinder grazing light scattering.
Therefore, this kind of {\it light-emitters} might not correlate with
any of the reciprocal versions (SPP flashlights) seen in PSTM
experiments.\cite{smol,bozh}
\subsection{Large width limit}
\label{sec_res_lwl}
Although the energy conservation criterion is reasonably well
satisfied in our calculations, even for defects wider than those used in
Fig.~\ref{fig_rts} (we have reached up to $a/\lambda=20$), one has to be
careful when interpreting the results in the limit
$a/\lambda\rightarrow\infty$.
It turns out that the determination
of the behavior of defects in this limit is important, since different
tendencies have been encountered for protuberances and indentations
(total radiation and transmission, respectively).
The analysis of the appropriate defect width that yields maximum coupling
is not simple even if making use of the Born approximation, since it
requires the evaluation of the integral of $\hat{s}(q-k^R)$ for all
homogeneous waves $|q|<\omega/c$. And yet such an approximation does not
properly describe the formally exact numerical calculations.
Alternatively, we have carried out an
analytic calculation based on the use of a boundary condition similar to
the Kirchhoff approximation. The approach relies on the expression for
the scattering amplitude in terms of an integral equation along the
surface with the magnetic field and its normal derivative inside the
integrand (cf. Refs. \onlinecite{josaa91,ann} for the integral equation
formulation, and Ref. \onlinecite{prb96} for its version making use of
the IBC on a flat surface). By assuming that the surface magnetic field
is given by the incoming SPP, the scattering amplitude reads in the large
width limit:
\begin{eqnarray}
\lim_{a/\lambda\rightarrow\infty} R(q,\omega)=
2\pi\delta\bbox(q-k^R(\omega)\bbox)(1+\pi s_0).
\end{eqnarray}
Although it does not satisfy energy conservation (not surprisingly, due to
the approximation involved), the former result gives an estimation of the
limiting behavior shown above (see Fig.~\ref{fig_rts}): SPP transmission
saturates for indentations ($s_0>0$), whereas protuberances ($s_0<0$)
tend to decrease SPP transmission (in agreement with the numerical
calculations, yet the exact limit is not predicted).
\section{CONCLUSIONS}
\label{sec_con}
We have presented a theoretical formulation that describes in a
rigorous manner the scattering of a surface plasmon polariton
propagating along a planar vacuum-metal interface by a
one-dimensional obstacle modeled through an impedance boundary
condition. By solving the $k$-space scattering integral equation
upon which the formulation is based, the angular spectrum of
the scattered electromagnetic field in the vacuum half-space
above the metal surface can be calculated, which in turn allows
us to obtain the near electric and magnetic fields, the amplitudes
of the reflected and transmitted SPP, and the angular distribution
of the intensity of radiated waves resulting from the conversion of
SPP into volume waves. A numerical method to solve the scattering
integral equation has been put forth.
We have made use of these calculation methods to study the SPP
scattering by one-dimensional Gaussian defects, either protuberances
or indentations. In particular, the dependence of the scattering
process on the surface defect parameters has been analyzed.
Several conclusions can be drawn from our results with respect to the
behavior of Gaussian protuberances or indentations.
SPP reflection is only significant for very narrow surface defects,
with half-widths $a<c/\omega$. Our near field results explicitly show
that in this case protuberances and indentations behave alike, the
latter reflecting SPP slightly more efficiently.
The dependence of the SPP reflection coefficient on the half-width
confirms for different defect heights the condition predicted in Ref.
\onlinecite{apl98} of maximum SPP reflection, leading to the {\it
plasmon mirrors} seen in PSTM experiments.\cite{smol}
For wider Gaussian defects, protuberances and indentations yield a
entirely different picture, the only common feature being the negligible
contribution to SPP reflection. Protuberances, on the one hand,
increasingly radiate more light at near grazing scattering angles at the
expense of SPP transmission. They behave as {\it light-emitters} with
coupling efficiencies approaching 100\% with increasing half-width. The
higher the defect is, the larger the SPP-light conversion.
On the other hand, indentations tend to total SPP transmission without
radiation with increasing half-width. The increase (decrease) of the SPP
transmission (light coupling) occurs in an oscillatory manner starting
from an absolute minimum (maximum) in transmission (radiation) for small
half-widths, the period of the oscillations being related to the defect
impedance in a way reminiscent of a cavity-like effect. Interestingly, we
have found that for sufficiently deep indentations, this maximum
radiation value can be extremely large (even larger than 90\%), so that
the Gaussian indentation thus behave as a light-emitter.
Our results and discussion provide a thorough picture of the different
aspects of SPP scattering by surface defects which, besides being
interesting in itself as a scattering process, appears to be useful in a
number of related problems.
\cite{ebbe98,schro98,nie97,klar98,smol95,smol,bozh,smol99,pasc98,apl98}
It is rigorous for one-dimensional defects and indeed sheds light on the
two-dimensional case, and can in turn explain and predict experimental
results.\cite{smol,bozh} In this regard, it would be interesting to
perform experiments on metal surfaces with defects of controled profile.
With respect to the 2D case, it should be emphasized that the recent work
by Shchegrov {\it et al.}\cite{snm97} for circularly symmetric surface
defects reproduces the radiation pattern with peculiar lobes in the
azimuthal angle dependence observed experimentally.\cite{smol}
However, further theoretical work is needed that could
address more complicated geometries used in the experiments and/or
unexplained processes involving surface plasmon-polaritons.
\cite{ebbe98,nie97,smol99}
\acknowledgments
This work was supported in part by Army Research Office Grant No.
DAAH 0-96-1-0187, and by both the Spanish DGES (Grants No.
PB96-0844-C02-02 and PB97-1221) and the Consejo Superior de
Investigaciones Cient{\'\i}ficas.
J.A.S.-G. acknowledges fruitful discussions with J. M. S{\'a}iz.
|
\section*{Introduction}
This paper is concerned with the local statistics of the simultaneous
zeros of $k$ random holomorphic sections $s_1, \dots, s_k \in H^0(M,
L^N)$ of the $N^{\rm th}$ power $L^N$
of a positive Hermitian holomorphic line bundle $(L,h)$ over a compact
K\"ahler\ manifold $M$ (where $k\le m=\dim M$). The terms `random' and
`statistics' are with respect to a natural Gaussian probability measure
$d \nu_N$ on
$H^0(M, L^N)$ which we define below. In the special case where
$M = \C\PP^m$ and $L$ is the
hyperplane section bundle $\mathcal{O}(1)$, sections of $L^N$ correspond
to holomorphic
polynomials of degree $N$, and $(H^0(\C\PP^m, \mathcal{O}(N)), d \nu_N)$
is known as the
ensemble of ${\operatorname{SU}}(m+1)$ polynomials in the physics literature.
To obtain local
statistics, we expand a ball $U$ around a given point $z^0$ by a factor
$\sqrt N$ so that the average density of simultaneous scaled
zeros is independent of $N$. We then ask whether the simultaneous
scaled zeros
behave as if thrown independently in $\sqrt{N} U$ or how they are
correlated.
Correlations between (unscaled) zeros are measured by the so-called
{\it $n$-point zero
correlation function\/} $K_{nk}^N(z^1,\dots,z^n)$, and those between
scaled zeros are
measured by the scaled correlation function $K_{nk}^N(\frac{z^1}{
\sqrt{N}},\dots,\frac{z^n}{ \sqrt{N}})$.
Our main result is that the large $N$
limits of the scaled
$n$-point correlation functions $K_{nk}^N(\frac{z^1}{
\sqrt{N}},\dots,\frac{z^n}{ \sqrt{N}})$ exist
and are universal, i.e. are independent of $M,\ L$ and $h$ as well as
the
point $z^0$. Moreover, the scaling limit correlation functions can be
calculated explicitly. We find that the limit correlations are short
range, i.e. that simultaneous scaled zeros behave quite independently
for large distances. On the other hand,
nearby zeros exhibit some degree of repulsion.
To state our problems and results more precisely, we begin with provisional
definitions of the correlation functions $K_{nk}^N(z^1,\dots,z^n)$ and of the
scaling limit. (See \S\S \ref{notation}--\ref{corfuns} for the complete
definitions and notation.) In order to provide a standard yardstick for our
universality results, we give $M$ the K\"ahler\ metric $\omega$ given by the
(positive) curvature form of $h$. The metrics $h$ and $\omega$ then induce a
Hilbert space inner product on the space $H^0(M, L^N)$ of holomorphic sections
of $L^N$, for each $N\ge 1$. In the spirit of \cite{SZ} we use this
$\mathcal{L}^2$-norm to define a Gaussian probability measure $d \nu_N$ on $H^0(M,
L^N)$. When we speak of a random section, we mean a section drawn at random
from this ensemble. More generally, we can draw $k$ sections $(s_1, \dots,
s_k)$ independently and at random from this ensemble. Let $Z_{(s_1, \dots,
s_k)}$ denote their simultaneous zero set and let $|Z_{(s_1, \dots, s_k)}|$
denote the ``delta measure" with support on $Z_{(s_1, \dots, s_k)}$ and with
density given by the natural Riemannian volume $(2m - 2k)$-form defined by the
metric $\omega$. To define the $n$-point zero correlation measure
$K_{nk}^N(z^1, \dots, z^n)$ we form the product measure
$$|Z_{(s_1, \dots, s_k)}|^n=\big(\underbrace{|Z_{(s_1, \dots,
s_k)}|\times\cdots\times|Z_{(s_1, \dots, s_k)}|}_{n}\big)\quad
\mbox{on}\quad
M^n:= \underbrace{M\times\cdots\times
M}_n\,.$$ To avoid trivial self-correlations, we puncture out the
generalized diagonal
in $M^n$ to get the punctured product space
$$M_n=\{(z^1,\dots,z^n)\in M^n: z^p\ne z^q \ \ {\rm for} \ p\ne q\}\,.$$
We then restrict $|Z_{(s_1, \dots, s_k)}|^n$ to $M_n$ and define
$K_{nk}^N(z^1, \dots, z^n)$
to be the
expected value $E(|Z_{(s_1, \dots, s_k)}|^n)$ of this measure with
respect to $\nu_N$. When $k =
m$, the simultaneous
zeros almost surely form a discrete set of points and so this case is
perhaps the most vivid. Roughly speaking,
$K_{nk}^N(z^1, \dots, z^n)$ gives the probability density
of finding
simultaneous zeros at
$(z^{1}, \dots, z^n)$.
The first correlation function $K_{1kN}$ just gives the expected
distribution of simultaneous zeros of $k$ sections.
In a previous paper
\cite{SZ} by two of the authors, it was shown (among other things) that
the expected
distribution of zeros is asymptotically uniform; i.e.
$$K_{1k}^N(z^0)=c_{mk}N^k+O(N^{k-1})\,,$$ for any positive line bundle
(see
\cite[Prop.~4.4]{SZ}). The question then arises of
determining the
higher correlation functions. As was first observed by \cite{BBL} and
\cite
{H} for ${\operatorname{SU}}(2)$ polynomials and by \cite{BD} for real polynomials in
one
variable, the zeros of a random polynomial are non-trivially correlated,
i.e. the zeros are not thrown down like independent points. We will
prove
the same for all ${\operatorname{SU}}(m+1)$ polynomials and hence, by universality of
the scaling limit, for any $M, L, h$.
To introduce the scaling limit, let us return to the case $k = m$ where
the simultaneous zeros form a discrete set of points. Since an
$m$-tuple of sections of $L^N$ will have $N^m$ times
as many zeros as $m$-tuples of sections of $L$, it is natural to
expand
$U$ by a factor of $\sqrt{N}$ to get a density of zeros that is
independent
of $N$. That is, we choose coordinates
$\{z_q\}$ for which
$z^0=0$ and
$\omega(z^0)=\frac{i}{2}\sum_q dz_q\wedge d\bar z_q$
and then rescale
$z \mapsto \frac{z}{
\sqrt{N}}$. Were the zeros thrown independently and at random on $U$, the
conditional probability density of finding a simultaneous zero at a
point
$w$ given a zero at $z$ would
be a constant independent of $(z,w)$. Non-trivial correlations
(for any codimension $k\in\{1,\dots,m\}$) are measured by the
difference between $1$ and the (normalized) {\it
$n$-point scaling limit zero correlation function\/}
$$\widetilde K_{nkm}^\infty(z^1,\dots,z^n)
=\lim_{N\to\infty}\left(c_{mk}N^k\right)^{-n}
{K_{nk}^N\left(\frac{z^1}{ \sqrt{N}},\dots,\frac{z^n}{
\sqrt{N}}\right)}\,,\qquad
(z^1,\dots,z^n)\in U_n\,.$$
Our main result (Theorem \ref{uslc}) is universality of the scaling
limit
correlation functions:
\bigskip\begin{quote}
{\it The $n$-point scaling limit zero correlation function
$\widetilde K_{nkm}^\infty(z^1,\dots,z^n)$ is given by a universal rational
function, homogeneous of degree $0$, in the values of the function
$e^{i\Im
(z\cdot \bar w)-{\frac{1}{2}} |z-w|^2}$ and its first and second derivatives at
the
points
$(z,w)=(z^p,z^{p'})$, $1\le p,p'\le n$. Alternately it is a
rational function in $z^p_q, \bar z^p_q, e^{z^p\cdot \bar z^{p'}}$}
\end{quote}
\bigskip The function $e^{i\Im (z\cdot \bar w)-{\frac{1}{2}} |z-w|^2}$ which
appears
in the universal scaling limit is (up to a constant factor) the Szeg\"o
kernel
$\Pi_1^\H(z,w)$ of level
one for the reduced Heisenberg group ${\bf H}^n_{\rm red}$ (cf.\
\S \ref{notation}). Its appearance here owes to the fact that the
correlation
functions can be expressed in terms of the Szeg\"o kernels $\Pi_N(x,y)$
of
$L^N$. I.e., let $X$ denote the circle bundle over $M$ consisting of
unit
vectors in
$L^*$; then $\Pi_N(x,y)$ is the kernel of the orthogonal projection
$\Pi_N:\mathcal{L}^2(X)\to \mathcal{H}^2_N(X)
\approx H^0(M, L^N)$. Indeed we have (Theorem \ref{npointcor}):
\bigskip
\begin{quote} {\it The $n$-point correlation $\widetilde
K_{nk}^N(z^1,\dots,z^n)$
is
given by the above universal
rational function, applied this time to the values of the Szeg\"o kernel
$\Pi_N$ and its first and second derivatives at the points
$(z^p,z^{p'})$. }
\end{quote}
\bigskip
In view of this relation between the correlation functions and the
Szeg\"o
kernel, it suffices for the proof of the universality theorem \ref{uslc}
to
determine the scaling limit of the Szeg\"o kernel $\Pi_N$ and to show
its
universality. Indeed we shall show (Theorem \ref{neardiag}) that:
\bigskip
\begin{quote} {\it Let $(z_1,\dots,z_m,\theta)$ denote local coordinates
in a neighborhood $\widetilde U\approx U\times S^1$ of a point $(z^0,\lambda)\in X$
(where $(z_1,\dots,z_m)$ are the above local coordinates about
$z_0\in M$). We then
have
$$N^{-m}\Pi_N\left(\frac{z}{\sqrt{N}},\frac{\theta}{N};
\frac{z'}{\sqrt{N}},\frac{\theta'}{N}\right)=
\Pi^\H_1(z,\theta;z',\theta') + O(N^{-1/2})\;.$$
}\end{quote}
\bigskip The fact that the correlation functions can be expressed
in terms of the Szeg\"o kernel may be explained in (at least) two
ways. The
first
is that the correlation functions may be expressed in terms of the
joint probability
density $D_{nk}^N(x,\xi;z)dxd\xi$ of the
(vector-valued) random variable
$$(x,\xi)=\big[x^p,\xi^p\big]_{1\le p\le n}\;,\qquad
x^p=(s_1(z^p),\dots, s_k(z^p)),\quad \xi^p=
(\nabla s_1(z^p),
\dots, \nabla s_k(z^p))$$ given by the values
of the $k$
sections and of their covariant derivatives at the $n$ points $\{z^p\}$.
Our method of computing the correlation functions is based on the
following
probabilistic formula (Theorem \ref{density}):
\bigskip
\begin{quote} {\it For $N$ sufficiently large so that the density
$D_{nk}^N
(x,\xi;z)$ is given by a continuous function, we have
$$
K_{nk}^N(z)=
\int d\xi\,D_{nk}^N(0,\xi;z)\prod_{p=1}^n
\det \left(\xi^{p}_j\xi^{p*}_{j'}\right)_{1\le j,j'\le k}\,,
\quad z=(z^1,\dots,z^n)\in M_n\,,$$
where $\xi=(\xi^1,\dots,\xi^n)$
and $\xi^{p*}_j:L^N_{z^p}\to T_{M,z^p}$ denotes
the adjoint to $\xi^p_j:T_{M,z^p}\to L^N_{z^p}$.}\end{quote}
\bigskip
This formula, which is valid in a more general setting, is based on
the approach of Kac \cite{Ka} and Rice \cite{Ri} (see also \cite {EK}) for
zeros of functions on ${\mathbb R}^1$, and of \cite{Halp} for zeros of (real) Gaussian
vector fields. Since our probability measure $d\nu_N$ (on the space of
sections) is Gaussian, it follows that $D_{nk}^N$ is also a Gaussian density.
It will be proved in \S \ref{D-meet-S} that the covariance matrix of this
Gaussian may be expressed entirely in terms of $\Pi_N$ and its covariant
derivatives. This type of formula for the correlation function of zeros was
previously used in \cite{BD}, \cite {H} and the works cited above.
We believe that this formula will have interesting applications in geometry.
A second
link between correlation functions and Szeg\"o kernels is given by the
Poincar\'e-Lelong
formula. In fact, this was our original approach to computing the
correlation
functions in the codimension 1 case. For the sake of brevity, we will
not
discuss this approach here; instead we refer the reader to our companion
article
\cite{BSZ}.
{From} the universality of our answers, it follows that the scaling
limit
pair
correlation functions depend only on the distance between points:
$$\widetilde K^\infty_{2km}(z^1,z^2)=\kappa_{km}(r)\,,\qquad r=|z^1-z^2|\,,$$
where $\kappa_{km}$ depends only on the dimension $m$ of $M$ and the
codimension $k$ of the zero set. In \S \ref{formulas}, we give explicit
formulas for the limit pair correlation functions $\kappa_{km}$
in some special cases.
Our calculation uses the Heisenberg model, which (although
noncompact) is the
most natural one since the scaled Szeg\"o kernels are all equal to
$\Pi_1$,
and there is no need in this case to take a limit. We also discuss the
hyperplane section bundle
$\mathcal{O} (1)
\to
\C\PP^m$, which is the most studied, since the sections of its powers are
the ${\operatorname{SU}}(m+1)$ polynomials---homogeneous polynomials in
$m+1$ variables---and the case $m=1$ (the ${\operatorname{SU}}(2)$ polynomials)
appears frequently in the physics literature (e.g., \cite{BBL, FH, H,
KMW,
PT}). We give expressions for the zero correlations $K^N_{nk}$ for the
${\operatorname{SU}}(m+1)$ polynomials and by letting $N\to\infty$, we obtain an
alternate derivation of our universal formula for the scaling limit
correlation.
We show (Theorem \ref{shortrange}) that $\kappa_{km}(r)=
1+O(r^4e^{-r^2})$ as $r\to +\infty$, and hence these correlations are
short
range in that they differ from the case of independent random points by
an
exponentially decaying term. We observe that when $\dim M=1$,
there is a strong repulsion between nearby zeros in the sense that
$\kappa_{11}(r) \to 0$ as $r \to 0$, as was noted by Hannay \cite{H} and
Bogomolny-Bohigas-Leboeuf
\cite{BBL} for the case of ${\operatorname{SU}}(2)$ polynomials. These asymptotics
are illustrated by the following graph (see also \cite{H}):
\vspace{-.5in}
\begin{figure}[ht]\label{kappa11}\centering
\epsfig{file=k11.eps,height=2.5in}
\caption{The 1-dimensional limit pair correlation function
$\kappa_{11}$}
\end{figure}
For $\dim M=2$, the simultaneous scaled
zeros of a random pair $(s_1, s_2)$ of sections
still exhibit a mild repulsion ($\lim_{r \to 0} \kappa_{22}(r)
=\frac{3}{4}$), as illustrated in Figure~2 below.
\vspace{-.1in}
\begin{figure}[ht]\label{kappa22}\centering
\epsfig{file=k22.eps,height=2.5in}
\caption{The limit pair correlation function $\kappa_{22}$}
\end{figure}
The function $\kappa_{mm}(r)$ can be interpreted as the normalized
conditional probability of finding a zero near a point $z^1$ given that
there
is a zero at a second point a scaled distance $r$ from $z^1$ (in the
case of discrete zeros in $m$ dimensions). The above graphs show that
for
dimensions 1 and 2, there is a unique scaled distance where this
probability
is maximized. It would be interesting to explore the dependence of the
correlations on the dimension. To ask one concrete question, do the
simultaneous scaled zeros in the point case become more and more
independent
in the sense that $\kappa_{mm} (r) \to 1$ as the dimension $m \to
\infty$?
When $k<m$, the zero sets are subvarieties of positive dimension $m-k$;
in
this case the expected volume of the zero set in a small spherical shell
of
radius $r$ and thickness $\varepsilon$ about a point in the zero set must be
$\sim \varepsilon
r^{2m-2k-1}$. Hence we have $\kappa_{km}(r)\sim r^{-2k}$, for small
$r$. The
graph of the limit correlation function for the case $m=2,k=1$ is given
in
Figure~3 below.
\vspace{-.25in}
\begin{figure}[ht]\label{kappa12}\centering
\epsfig{file=k12.eps,height=2.5in}
\caption{The limit pair correlation function $\kappa_{12}$}
\end{figure}
To end this introduction, we would like to link our methods and results
at least
heuristically to a long tradition of (largely heuristic) results on
universality and scaling in statistical mechanics (cf. \cite{FFS}). One
may
view the rescaling
transformation on $U$
as generating a renormalization group. The intuitive picture in
statistical
mechanics is that the renormalization
group should carry a given system (read ``$L \to M$'') to the fixed
point
of the renormalization
group, i.e. to the scale invariant situation. We observe that the
local rescaling
of $U$ is nothing other than the Heisenberg dilations
$\delta_{\sqrt{N}}$ on
${\bf H}^m_{\rm red}$. Since these dilations are automorphisms of the
(unreduced) Heisenberg group,
the Szeg\"o kernel of ${\bf H}^m$ is invariant under these dilations;
i.e., it is the fixed point of the renormalization group. As predicted
by
this intuitive picture, we find that in the scaling limit all the
invariants of the line
bundle, in particular its zero-point
correlation functions, are drawn to their values for the fixed point
system
(read ``Heisenberg model'').
\bigskip
\section{Notation}\label{notation}
We begin with some notation and basic properties of sections of
holomorphic
line bundles, their zero sets, Szeg\"o kernels, and Gaussian measures.
We
also provide two examples that will serve as model cases for studying
correlations of zeros of sections of line bundles in the high power
limit.
\subsection{Sections of holomorphic line bundles}\label{cxgeom}
In this section, we introduce the basic complex analytic objects:
holomorphic
sections and the currents of integration over their zero sets. We also
introduce Gaussian probability measures on spaces of holomorphic
sections.
For background in complex geometry,
we refer to \cite{GH}.
Let $M$ be a compact complex manifold and let $L\to M$ be a holomorphic line
bundle with a smooth Hermitian metric $h$; its curvature 2-form $\Theta_h$ is
given locally by
\begin{equation}\label{curvature}\Theta_h=-\partial\dbar
\log\|e_L\|_h^2\;,\end{equation} where $e_L$ denotes a local holomorphic
frame
(= nonvanishing section) of $L$ over an open set $U\subset M$, and
$\|e_L\|_h=h(e_L,e_L)^{1/2}$ denotes the $h$-norm of $e_L$.
We say that $(L,h)$ is positive if the (real) 2-form
$\omega=\frac{\sqrt{-1}}{2}\Theta_h$ is positive, i.e., if $\omega$ is a K\"ahler\
form. We henceforth assume that $(L,h)$ is positive, and
we give $M$
the
Hermitian metric corresponding to the K\"ahler\ form
$\omega$ and the induced Riemannian volume
form
\begin{equation}\label{dV} dV_M= \frac{1}{m!}
\omega^m\;.\end{equation}
Since $\frac{1}{\pi}\omega$ is a
de Rham representative of the Chern class $c_1(L)\in H^2(M,{\mathbb R})$, the
volume of
$M$ equals $\frac{\pi^m}{m!}c_1(L)^m$.
The space $H^0(M, L^{N})$ of global holomorphic sections of
$L^N=L\otimes\cdots\otimes L$ is a finite dimensional complex vector space.
(Its dimension, given by the Riemann-Roch formula for large $N$, grows like
$N^m$. By the Kodaira embedding theorem, the global sections of $L^N$ give an
embedding into a projective space for $N\gg 0$, and hence $M$ is a {\it
projective algebraic manifold.\/}) The metric $h$ induces Hermitian metrics
$h^N$ on $L^N$ given by $\|s^{\otimes N}\|_{h^N}=\|s\|_h^N$. We give
$H^0(M,L^N)$ the Hermitian inner product
\begin{equation}\label{inner}\langle s_1, s_2 \rangle = \int_M h^N(s_1,
s_2)dV_M \quad\quad (s_1, s_2 \in H^0(M,L^N)\,)\;,\end{equation} and we
write
$|s|=\langle s,s \rangle^{1/2}$.
We now
explain our concept of a ``random section." We are
interested in expected values and correlations of zero sets of
$k$-tuples of
holomorphic sections of powers $L^N$. Since the zeros do not depend on
constant factors, we could suppose our sections lie in the unit
sphere in
$H^0(M,L^N)$ with respect to the Hermitian inner product (\ref{inner}),
and we pick random sections with respect to the spherical measure.
Equivalently, we could suppose that $s$ is a random element of the
projectivization
${\mathbb P}
H^0(M,L^N)$. Another equivalent approach is to use Gaussian measures on
the
entire space $H^0(M,L^N)$. We shall use the third approach, since
Gaussian
measures seem the best for calculations. Precisely, we give
$H^0(M,L^N)$ the complex
Gaussian probability measure
\begin{equation}\label{gaussian}d\nu_N(s)=\frac{1}{\pi^m}e^
{-|c|^2}dc\,,\qquad s=\sum_{j=1}^{d_N}c_jS_j^N\,,\end{equation} where
$\{S_j^N\}$ is an orthonormal basis for
$H^0(M,L^N)$ and $dc$ is $2d_N$-dimensional Lebesgue measure. This
Gaussian
is characterized by the property that the $2d_N$ real variables $\Re
c_j,
\Im c_j$ ($j=1,\dots,d_N$) are independent random variables with mean 0
and
variance ${\frac{1}{2}}$; i.e.,
$${\mathbf E}\, c_j = 0,\quad {\mathbf E}\, c_j c_k = 0,\quad {\mathbf E}\, c_j \bar c_k = \delta_{jk}\,.$$
Here and throughout this paper, ${\mathbf E}\,$ denotes expectation.
In general, a {\it complex Gaussian measure\/} (with mean 0) on a finite
dimensional complex vector space $V$ is a measure $\nu$ of the form
(\ref{gaussian}), where the $c_j$ are the coordinates with respect to
some
basis. Explicitly, the complex Gaussian measures on ${\mathbb C}^m$ are the
probability measures of the form
\begin{equation}\label{gaussian2} \frac{e^{-\langle
\Delta^{-1}z,z\rangle}}{\pi^m\det
\Delta}dz \end{equation} where $\Delta=(\Delta^j_k)$ is a
positive definite Hermitian matrix and
$$\langle \zeta, z\rangle=\zeta\cdot\bar z=\sum_{q=1}^m \zeta_q\bar
z_q$$
denotes the standard Hermitian inner product in ${\mathbb C}^m$. For the
Gaussian
measure (\ref{gaussian2}), we have
\begin{equation}\label{gaussian3} {\mathbf E}\,(z_j z_k)=0, \qquad {\mathbf E}\,(z_j \bar
z_k)=\Delta^j_k\,.\end{equation}
If $\nu$ is a complex Gaussian on $V$ and
$\tau:V\to
\widetilde V$ is a surjective linear transformation, then $\tau_*\nu$ is a
complex
Gaussian on $\widetilde V$. In particular, if $\widetilde V={\mathbb C}^m$, then, $\tau_*\nu$
is of
the form (\ref{gaussian2}), where the covariance matrix $\Delta$ is given
by
(\ref{gaussian3}) with $z_j=z_j\circ\tau:V\to{\mathbb C}$.
We shall consider the space $\mathcal{S}=H^0(M,L^N)^k$ ($1\le k \le m$) with
the
probability measure $d\mu=d \nu\times\cdots\times d\nu$, which is also
Gaussian. Picking a random element of $\mathcal{S}$ means picking $k$
sections of
$H^0(M,L^N)$ independently and at random. For
$s=(s_1,\dots,s_k)\in\mathcal{S}$, we
let
$$Z_s=\{z\in M: s_1(z)=\cdots=s_k(z)=0\}$$
denote the zero set of $s$.
Note that if $N$ is sufficiently large so that $L^N$ is base point free,
then for
$\mu$-a.a.
$s\in\mathcal{S}$, we have
${\operatorname{codim\,}} Z_s=k$. (Indeed, the set of $s$
where
${\operatorname{codim\,}} Z_s<k$ is a proper algebraic subvariety of $H^0(M,L^N)^k$. In
fact, by Bertini's theorem, the $Z_s$ are smooth submanifolds of complex
dimension $m-k$ for almost all $s$, provided $N$ is large enough so that
the
global sections of $L^N$ give a projective embedding of $M$, but we do
not
need this fact here.) For these $s$, we let
$|Z_s|$ denote Riemannian
$(2m-2k)$-volume along the regular points of $Z_s$, regarded as a
measure on
$M$:
\begin{equation}\label{volmeasure} (|Z_s|,\phi)=\int_{Z_s^{\rm reg}}\phi
d{\operatorname{Vol}}_{2m-2k}=\frac{1}{(m-k)!}\int_{Z_s^{\rm reg}}\phi
\omega^{m-k}\,.\end{equation} It was shown by Lelong \cite{Le}
(see also \cite{GH}) that
the integral in (\ref{volmeasure}) converges.
(In fact, $|Z_s|$ can be regarded as the total variation measure of
the closed current of integration
over $Z_s$.)
We regard $|Z_s|$ as a measure-valued random variable on the
probability space $(\mathcal{S},d\mu)$; i.e., for each test function
$\phi\in\mathcal{C}^0(M)$, $(|Z_s|,\phi)$ is a complex-valued random variable.
\subsection{Szeg\"o kernels}\label{Szegokernels}
As in \cite{Z,SZ} we now lift the analysis of holomorphic sections
over $M$ to
a certain $S^1$ bundle $X \to M$. This is a useful approach to the
asymptotics of powers
of line bundles and goes back at least to \cite{BG}.
We let $L^*$ denote the dual line bundle to $L$, and we consider the
circle
bundle $X=\{\lambda \in L^* : \|\lambda\|_{h^*}= 1\}$, where $h^*$ is the norm
on
$L^*$ dual to $h$. Let $\pi:X\to M$ denote the bundle map; if $v\in
L_z$, then
$\|v\|_h=|(\lambda,v)|$, $\lambda\in X_z=\pi^{-1}(z)$. Note that $X$ is the
boundary
of the
disc bundle $D = \{\lambda \in L^* : \rho(\lambda)>0\}$, where
$\rho(\lambda)=1-\|\lambda\|^2_{h^*}$. The disc bundle $D$ is strictly
pseudoconvex in
$L^*$, since $\Theta_h$ is positive, and hence $X$ inherits the
structure of
a strictly pseudoconvex CR manifold. Associated to $X$ is the contact
form
$\alpha= -i\partial\rho|_X=i\bar\partial\rho|_X$. We also give $X$ the volume
form
\begin{equation}\label{dvx}dV_X=\frac{1}{m!}\alpha\wedge
(d\alpha)^m=\alpha\wedge\pi^*dV_M\,.\end{equation}
The setting for our analysis of the Szeg\"o kernel is the Hardy space
$\mathcal{H}^2(X)
\subset \mathcal{L}^2(X)$ of square-integrable CR functions on $X$, i.e.,
functions that are annihilated by the
Cauchy-Riemann operator $\bar\partial_b$ (see \cite[pp.~592--594]{Stein}) and
are
$\mathcal{L}^2$ with respect to the inner product
\begin{equation}\label{unitary} \langle F_1, F_2\rangle
=\frac{1}{2\pi}\int_X
F_1\overline{F_2}dV_X\,,\quad F_1,F_2\in\mathcal{L}^2(X)\,.\end{equation}
Equivalently, $\mathcal{H}^2(X)$
is the space of boundary values of holomorphic functions on $D$ that
are
in
$\mathcal{L}^2(X)$. We let $r_{\theta}x =e^{i\theta} x$ ($x\in X$) denote the
$S^1$
action on $X$ and denote its infinitesimal generator by
$\frac{\partial}{\partial \theta}$. The $S^1$ action on $X$ commutes
with $\bar{\partial}_b$; hence $\mathcal{H}^2(X) = \bigoplus_{N
=0}^{\infty} \mathcal{H}^2_N(X)$ where $\mathcal{H}^2_N(X) =
\{ F \in \mathcal{H}^2(X): F(r_{\theta}x)
= e^{i
N \theta} F(x) \}$. A section $s$ of $L$ determines an equivariant
function
$\hat{s}$ on $L^*$ by the rule $\hat{s}(\lambda) = \left(\lambda,
s(z)
\right)$ ($\lambda \in L^*_z, z \in M$). It is clear that if $\tau \in
{\mathbb C}$
then $\hat{s}(z, \tau \lambda) = \tau \hat{s}$. We henceforth restrict
$\hat{s}$ to $X$ and then the equivariance property takes the form
$\hat{s}(r_{\theta} x) = e^{i \theta}\hat{s}(x)$. Similarly, a section
$s_N$
of $L^{N}$ determines an equivariant function $\hat{s}_N$ on $X$: put
$$\hat{s}_N(\lambda) = \left( \lambda^{\otimes N}, s_N(z)
\right)\,,\quad
\lambda\in X_z\,,$$ where $\lambda^{\otimes N} = \lambda \otimes
\cdots\otimes
\lambda$;
then $\hat s_N(r_\theta x) = e^{iN\theta} \hat s_N(x)$. The map
$s\mapsto
\hat{s}$ is a unitary equivalence between $H^0(M, L^{ N})$ and
$\mathcal{H}^2_N(X)$. (This follows from (\ref{dvx})--(\ref{unitary}) and the
fact
that
$\alpha= d\theta$ along the fibers of $\pi:X\to M$.)
We let $\Pi_N : \mathcal{L}^2(X) \rightarrow \mathcal{H}^2_N(X)$ denote the
orthogonal
projection. The Szeg\"o kernel $\Pi_N(x,y)$ is defined by
\begin{equation} \Pi_N F(x) = \int_X \Pi_N(x,y) F(y) dV_X (y)\,,
\quad F\in\mathcal{L}^2(X)\,.
\end{equation} It can be given as
\begin{equation}\label{szego}\Pi_N(x,y)=\sum_{j=1}^{d_N}\widehat
S_j^N(x)\overline{\widehat S_j^N(y)}\,,\end{equation} where
$S_1^N,\dots,S_{d_N}^N$ form an orthonormal basis of $H^0(M,L^N)$. Pick
a
local holomorphic frame $e_L$ for $L$ over an open subset $U\subset M$,
let
$e_L^*$ denote the dual frame, and write
$h(z)=h(e_L(z),e_L(z))=\|e_L\|_h^2$.
The map
$(z,e^{i\theta})\mapsto e^{i\theta}h(z)^{1/2} e_L^*(z)$ gives an
isomorphism
$U\times S^1\approx \pi^{-1}(U)\subset X$, and we use the coordinates
$(z,\theta)$ to identify points of $\pi^{-1}(U)$. For $s\in H^0(M,L^N)$,
we have
\begin{equation} \label{hats}\hat s(z,\theta)=
\left\langle s(z),e^{iN\theta}h(z)^{N/2}e^*_L
(z)\right\rangle = e^{iN\theta}h(z)^{N/2} f(z),\quad s=fe_L^{\otimes
N}\,.
\end{equation}
Although the Szeg\"o kernel is defined on $X$, its absolute value is
well-defined on $M$ as follows: writing $S_j^N=f^N_j e_L^{\otimes N}$,
we
have
\begin{equation}\label{PiN}\Pi_N(z,\theta;w,\phi)
=e^{iN(\theta-\phi)}\Pi_N(z,0;w,0)= e^{iN(\theta-\phi)}h(z)^{N/2}
h(w)^{N/2}\sum_{j=1}^{d_N}f^N_j(z)\overline{f^N_j(w)}\,,\end{equation}
for
$z,w\in U$. (Here we may take $U$ to be the disjoint union of connected
neighborhoods of $z$ and $w$, if $z$ is not close to $w$.) Thus we can
write
$$|\Pi_N(z,w)|= |\Pi_N(z,0;w,0)|\,,$$ which is independent of the choice
of
local frame $e_L$. On the diagonal we have $$\Pi_N(z,z)=
\Pi_N(z,\theta;z,\theta)=\sum_{j=1}^{d_N}\|S_j^N(z)\|_{h^N}\,.$$
The
Hermitian
connection $\nabla$ on $L$ induces the decomposition $T_X=T_X^{H}\oplus
T_X^{V}$ into horizontal and vertical components, and
we let $t^H$ denote the horizontal lift (to $X$) of a vector field $t$
in $M$.
We consider the horizontal operators on $X$:
$$d_{z_q}^H \buildrel {\operatorname{def}}\over = d_{(\d /\d z_q)^H}\,,\quad
d_{\bar z_q}^H\buildrel {\operatorname{def}}\over = d_{(\d /\d \bar z_q)^H}\,,$$
where $z_1,\dots,z_m,\theta$ denote local coordinates on
$X$. We note that
\begin{equation} \label{horizontal} d_{z_q}^H \hat s = (\nabla^N_{z_q}
s)\raisebox{2pt}{$\wh{\ }$}
\;,\quad s\in H^0(M,L^N)\,,\end{equation}
where $\nabla^N$ is the induced connection on $L^N$. We then have
\begin{eqnarray}d^H_{z_q}\Pi_N(z,\theta;w,\phi)
&=&\sum_{j=1}^{d_N}\left(\nabla^N_{z_q}
S_j^N\right)\raisebox{2pt}{$\wh{\ }$}(x)\overline{\widehat S_j^N(y)}\nonumber\\
&=&e^{iN(\theta-\phi)}h(z)^{N/2}
h(w)^{N/2}\sum_{j=1}^{d_N}f^N_{j;q}(z)\overline{f^N_j(w)}\,,\nonumber\\
\label{dPiN} d^H_{z_p} d^H_{\bar w_q}\Pi_N(z,\theta;w,\phi)&=&
\sum_{j=1}^{d_N}\left(\nabla^N_{z_p}
S_j^N\right)\raisebox{2pt}{$\wh{\ }$}(x)\overline{\left(\nabla^N_{w_q}
S_j^N\right)\raisebox{2pt}{$\wh{\ }$}(y)}\\
&=&e^{iN(\theta-\phi)}h(z)^{N/2}
h(w)^{N/2}\sum_{j=1}^{d_N}f^N_{j;p}(z)\overline{f^N_{j;q}(w)}\,,
\nonumber
\end{eqnarray} $$ \qquad \nabla^N_{z_q}=\nabla^N_{\d /\d z_q}\,,\quad
f^N_{j;q} =\frac{\d f}{\d z_q}+Nf^N_jh^{-1}\frac{\d
h}{\d z_q}\,.$$
We can also use (\ref{hats}) and (\ref{horizontal}) to describe the
horizontal
lift in local coordinates:
\begin{equation}\label{horizontal2} d^H_{z_q}= \frac{\d}{\d z_q}
-\frac{i}{2}
\frac{\d\log h}{\d z_q}\frac{\d}{\d\theta}\,.\end{equation}
\subsection{Model examples} In two special cases we can work out the
Szeg\"o
kernels and their derivatives
explicitly, namely for the hyperplane section bundle over $\C\PP^m$ and
for
the Heisenberg bundle over ${\mathbb C}^m$, i.e.
the trivial line bundle with curvature equal to the standard symplectic
form on ${\mathbb C}^m$.
These cases will be
important after we have proven universality, since scaling limits of
correlation functions
for all line bundles coincide with those of the model cases.
In fact, the two models are locally equivalent in the CR sense.
In the case of
$\C\PP^m$, the circle bundle $X$ is the $2m + 1$ sphere $S^{2m +
1}$, which is the boundary of the unit ball $B^{2m + 2} \subset
{\mathbb C}^{m+1}$. In the case of ${\mathbb C}^m$, the
circle bundle is the reduced Heisenberg group ${\bf H}^m_{\rm red}$,
which
is a discrete quotient of the simply connected
Heisenberg group ${\mathbb C}^m \times {\mathbb R}$. As is well-known, the latter is
equivalent (in the CR and contact sense) to the
boundary of $B^{2m + 2} $ (\cite{Stein}).
\subsubsection{${\operatorname{SU}}(m+1)$-polynomials}\label{example1}
For our first example, we let
$M=\C\PP^m$ and take
$L$ to be the hyperplane section bundle $\mathcal{O}(1)$.
Sections $s\in
H^0({\mathbb C}{\mathbb P}^m,\mathcal{O}(1))$ are linear functions on
${\mathbb C}^{m+1}$; the zero divisors $Z_s$ are projective hyperplanes. The line
bundle $\mathcal{O}(1)$ carries a natural metric $h_{{\operatorname{FS}}}$ given by
\begin{equation}\label{hfs} \|s\|_{h_{{\operatorname{FS}}}}([w])=\frac{|(s,w)|}{|w|}\;,
\quad\quad
w=(w_0,\dots,w_m)\in{\mathbb C}^{m+1}\;,\end{equation} for $s\in{\mathbb C}^{m+1*}\equiv
H^0({\mathbb C}{\mathbb P}^m,\mathcal{O}(1))$, where $|w|^2=\sum_{j=0}^m |w_j|^2$ and
$[w]\in{\mathbb C}{\mathbb P}^m$
is the complex line through $w$. The K\"ahler\ form on $\C\PP^m$ is the
Fubini-Study form
\begin{equation}
\omega_{{\operatorname{FS}}}=\frac{\sqrt{-1}}{2}\Theta_{h_{{\operatorname{FS}}}}=\frac{\sqrt{-1}}{2}
\partial\dbar \log |w|^2 \,.\end{equation}
The dual bundle $L^*=\mathcal{O}(-1)$ is the affine space
${\mathbb C}^{m+1}$ with the origin blown up, and $X=S^{2m+1}\subset{\mathbb C}^{m+1}$.
The
$N$-th tensor power of $\mathcal{O}(1)$ is denoted $\mathcal{O}(N)$. Elements
$s_N\in
H^0({\mathbb C}{\mathbb P}^m,\mathcal{O}(N))$ are homogeneous polynomials on ${\mathbb C}^{m+1}$ of
degree
$N$, and $\hat s_N=s_N|_{S^{2m-1}}$. The monomials
\begin{equation}\label{orthonormal}s^N_J =
\left[\frac{(N+m)!}{\pi^m j_0!\cdots j_m!}\right]^{\frac{1}{2}} z^J\,,\quad z^J=
z_0^{j_0}\cdots z_m^{j_m},\quad\quad J=(j_0,\ldots,j_m),\
|J|=N\end{equation}
form an orthonormal basis for $H^0({\mathbb C}{\mathbb P}^m,\mathcal{O}(N))$. (See \cite[\S
4.2]{SZ};
the extra factor $\left(\frac{m!}{\pi^m}\right)^{1/2}$ in
(\ref{orthonormal})
comes from the fact that here $\C\PP^m$ has the usual volume
$\frac{\pi^m}{m!}$, whereas in \cite{SZ}, the volume of $\C\PP^m$
is normalized to be 1.) Hence the Szeg\"o
kernel for $\mathcal{O}(N)$ is given by
\begin{equation}\label{Szegosphere} \Pi_N(x,y)=\sum_J
\frac{(N+m)!}{\pi^mj_0!\cdots j_m!}x^J \bar y^J = \frac{(N+m)!}
{\pi^mN!}\langle x,y\rangle^N\,.\end{equation}
Note that
$$\Pi(x,y)=\sum_{N=1}^\infty\Pi_N(x,y)=\frac{m!}{\pi^m}
(1-\langle x, y\rangle)^{-(m+1)} =
2\pi\times [\mbox{classical Szeg\"o kernel on} \
S^{2m+1}]\,.$$ (The factor
$2\pi$ is due to our normalization (\ref{unitary}).)
\subsubsection{The Heisenberg model}\label{example2}
Our second
example is the linear model ${\mathbb C}^m
\times {\mathbb C} \to {\mathbb C}^m$ for positive line bundles $L \to M$ over K\"ahler\
manifolds and their associated Szeg\"o kernels. It is most illuminating
to consider the associated
principal $S^1$ bundle ${\mathbb C}^m \times S^1 \to {\mathbb C}^m$, which may be
identified with the boundary of the disc bundle $D \subset L^*$ in the
dual line bundle. This $S^1$ bundle is the {\it
reduced Heisenberg group} ${\bf H}_{\rm red}^m$ (cf. \cite{F}, p. 23).
Let us recall its definition and properties. We start with the usual
(simply connected) Heisenberg group ${\bf
H}^m$ (cf. \cite{F} \cite{Stein}; note that different authors differ by
factors of $2$ and $\pi$
in various definitions). It is the
group ${\mathbb C}^m \times {\mathbb R}$ with group law
$$(\zeta, t) \cdot (\eta, s) = (\zeta + \eta, t + s + \Im (\zeta
\cdot
\bar{\eta})).$$
The identity element is $(0, 0)$ and $(\zeta, t)^{-1} = (- \zeta, - t)$.
Abstractly, the Lie algebra of ${\bf H}_m$ is spanned by elements $Z_1,
\dots, Z_m, \bar{Z}_1, \dots, \bar{Z}_m, T$ satisfying the canonical
commutation relations $[Z_j, \bar{Z}_k] = -i \delta_{jk} T$ (all other
brackets zero). Below
we will select such a basis of left invariant vector fields.
${\bf H}^m$ is a strictly convex CR manifold
which may be embedded in ${\mathbb C}^{m + 1}$ as the boundary of a strictly
pseudoconvex domain, namely
the upper half space $ \mathcal{U} ^m := \{z \in {\mathbb C}^{m + 1}: \Im z_{m + 1}
> {\frac{1}{2}} \sum_{j = 1}^m |z_j|^2 \}$.
The boundary of $\mathcal{U} ^m$ equals $\partial \mathcal{U} ^m = \{z \in {\mathbb C}^{m
+ 1}: \Im z_{m + 1} = {\frac{1}{2}}
\sum_{j = 1}^m |z_j|^2\}$.
${\bf H}^m$ acts simply transitively on $\partial \mathcal{U} ^m$ (cf.
\cite{Stein}, XII), and we get
an identification of ${\bf H}^m$ with $\partial \mathcal{U} ^m$ by:
$$[\zeta, t] \to (\zeta, t + i |\zeta|^2) \in \partial \mathcal{U} ^m.$$
The Szeg\"o projector of ${\bf H}^m$ is the operator $\Pi: \mathcal{L}^2({\bf
H}^m) \to
\mathcal{H}^2({\bf H}^m)$
of orthogonal projection onto boundary values of holomorphic functions
on
$\mathcal{U} ^m$ which lie in $\mathcal{L}^2$. The kernel of
$\Pi$ is given by (cf. \cite{Stein}, XII \S 2 (29))
$$\Pi(x,y) = K(y^{-1} x), \;\;\;\;\;\;\; K(x) = - C_m
\frac{\partial}{\partial t} [t + i |\zeta|^2]^{-m} \in\mathcal{D}'(\H^m)\,.$$
The linear model for the principal $S^1$ bundle described in \S 1.2 is
the
so-called reduced Heisenberg group ${\bf H}^m_{\rm red}={\bf H}^m/ \{(0,
2\pi
k): k \in {\mathbb Z}\} = {\mathbb C}^m \times S^1$ with group law $$(\zeta, e^{i t})
\cdot
(\eta, e^{i s}) = (\zeta + \eta, e^{i(t + s + \Im (\zeta \cdot
\bar{\eta}))}).$$ It is the principal $S^1$ bundle over ${\mathbb C}^m$
associated to
the line bundle $L_\H={\mathbb C}^m\times{\mathbb C}$. The metric on $L_\H$ with
curvature
$\Theta=\partial\dbar |z|^2$ is given by setting $h_\H(z)=e^{-|z|^2}$; i.e.,
$|f|_{h_\H}= |f|e^{-|z|^2/2}$. The reduced Heisenberg group ${\bf
H}^m_{\rm
red}$ may be viewed as the boundary of the dual disc bundle $D \subset
L^*_\H$
and hence is a strictly pseudoconvex CR manifold.
It seems most natural to approach the analysis of the Szeg\"o kernels on
${\bf
H}^m_{\rm red}$ from the representation-theoretic point of view. Let us
begin
with the case $N = 1$. We thus consider the space $\mathcal{V}_1
\subset
\mathcal{L}^2({\bf H}^m_{\rm red})$ of functions $f$ satisfying
$\frac{1}{i}\frac{\partial}{ \partial \theta} f = f$, which forms a
(reducible) representation of ${\bf H}^m_{\rm red}$ with central
character
$e^{i \theta}$. By the Stone-von Neumann theorem there exists a unique
(up to
equivalence) representation $(V_1, \rho_1)$ with this character and by
the
Plancherel theorem, $\mathcal{V}_1 \cong V_1 \otimes V_1^*$.
The space of CR functions in $\mathcal{V}_1$ is an irreducible invariant
subspace. Here, by CR functions we mean the functions satisfying the
left-invariant Cauchy-Riemann equations $\bar{Z}^L_q f = 0$ on ${\bf
H}^m_{\rm
red}$. Here, $\{\bar{Z}^L_q\}$ denotes a basis of the left-invariant
anti-holomorphic vector fields on ${\bf H}^m_{\rm red}$. Let us recall
their
definition: we first equip ${\bf H}^m_{\rm red}$ with its left-invariant
contact form $\alpha^L = \sum_q(u_qdv_q-v_qdu_q) +
d\theta$ ($\zeta=u+iv$). The left-invariant CR
holomorphic
(resp.\ anti-holomorphic) vector fields $Z_q^L$ (resp.\ $ \bar{Z}_q^L$)
are the
horizontal lifts of the vector fields $\frac{\partial}{\partial z_q}$
(resp.\
$\frac{\partial}{\partial \bar{z}_q}$) with respect to $\alpha^L$. They span
the
left-invariant CR structure of ${\bf H}^m_{\rm red}$ and the $Z_q^L$
obviously
have the form $Z_q^L = \frac{\partial}{\partial z_q} + A
\frac{\partial}{\partial \theta}$ where the coefficient $A$ is
determined by
the condition $\alpha^L(Z_q^L ) = 0$. An easy calculation gives:
$$Z_q^L = \frac{\partial}{\partial z_q} + \frac{i}{2}\bar{z}_q
\frac{\partial}{\partial \theta}, \;\;\;\; \bar{Z}^L_q =
\frac{\partial}{\partial \bar{z}_q} - \frac{i}{2} z_q
\frac{\partial}{\partial
\theta}.$$ The vector fields $\{\frac{\partial}{\partial \theta}, Z_q^L,
\bar{Z}_q^L\}$ span the Lie algebra of ${\bf H}^m_{\rm red}$ and satisfy
the
canonical commutation relations above.
We then define the Hardy space $\mathcal{H}^2 ({\bf H}^m_{\rm red})$ of
CR
holomorphic functions, i.e. solutions of $\bar{Z}_q^L f = 0$, which lie
in
$\mathcal{L}^2({\bf H}^m_{\rm red})$. We also put $\mathcal{H}^2_1 =
\mathcal{V}_1
\cap \mathcal{H}^2({\bf H}^m_{\rm red}).$ The group ${\bf H}^m_{\rm
red}$ acts
by left translation on $\mathcal{H}^2_1$. The generators of this
representation are the right-invariant vector fields $Z_q^R,
\bar{Z}_q^R$
together with $\frac{\partial}{\partial \theta}$. They are horizontal
with
respect to the right-invariant contact form $\alpha^R =
\sum_q(u_qdv_q-v_qdu_q) - d\theta$ and are given by:
$$Z_q^R = \frac{\partial}{\partial z_q} - \frac{i}{2}\bar{z}_q
\frac{\partial}{\partial \theta}, \;\;\;\; \bar{Z}^R_q =
\frac{\partial}{\partial \bar{z}_q} + \frac{i}{2} z_q
\frac{\partial}{\partial
\theta}\,.$$ In physics terminology, $Z_q^R$ is known as an annihilation
operator and $\bar{Z}_q^R$ is a creation operator.
The representation $\mathcal{H}^2_1$ is irreducible and may be identified with
the Bargmann-Fock space of entire holomorphic functions on ${\mathbb C}^m$ which are
square integrable relative to $e^{-|z|^2}$ (or equivalently, holomorphic
sections of the trivial line bundle $L_\H={\mathbb C}^m\times {\mathbb C}$ mentioned above, with
hermitian metric $h_\H=e^{-|z|^2}$). The identification goes as follows: the
function $\phi_0(z, \theta) := e^{i \theta} e^{-|z|^2/2}$ is CR holomorphic
and is also the ground state for the right invariant ``annihilation operator;''
i.e., it satisfies
$$\bar{Z}^L_q \phi_0(z, \theta) = 0 = Z_q^R \phi_0(z, \theta)\,.$$ Any element
$F(z,\theta)$ of $\mathcal{ H}^2_1$ may be written in the form $F (z, \theta)
= f(z) \phi_0.$ Then $\bar{Z}^L_q F =(\frac{\partial}{\partial \bar{z}_q} f )
\phi_0$, so that $F$ is CR if and only if $f$ is holomorphic. Moreover, $F \in
\mathcal{L}^2({\bf H}^m_{\rm red})$ if and only if $f$ is square integrable
relative to $e^{-|z|^2}$.
The Szeg\"o kernel $\Pi_1^\H(z, \theta, w, \phi)$ of ${\bf H}^m_{\rm
red}$ is by
definition the orthogonal projection from $\mathcal{L}({\bf H}^m_{\rm red})$
to
$H^2_1.$ As will be seen below, $\Pi_1^\H(z, \theta, w, \phi) =
\frac{1}{\pi_m}e^{i (\theta
-\phi )} e^{ (z\cdot\bar w -{\frac{1}{2}} |z|^2 -{\frac{1}{2}}|w|^2) }$, which is the
left
translate of $\phi_0$ by $(-w, - \phi)$. In the physics terminology it
is the
coherent state associated to the phase space point $w.$
So far we have set $N = 1$, but the story is very similar for any $N$.
We
define $\mathcal{H}^2_N$ as the space of square- integrable CR functions
transforming by $e^{i N \theta}$ under the central $S^1$. By the
Stone-von
Neumann theorem there is a unique irreducible $V_N$ with this central
character. The main difference to the case $N = 1$ is that $\mathcal{H}^2_N$
is of
multiplicity $N^m$. The Szeg\"o kernel $\Pi_N^\H(x,y)$ is the
orthogonal
projection to $\mathcal{H}^2_N$ and is given by the dilate of $\Pi^\H_1$.
Thus,
$$\Pi_N^\H(x,y) =\frac{1}{\pi^m} N^m e^{i N (\theta -\phi )} e^{
N(z\cdot\bar
w -{\frac{1}{2}} |z|^2 -{\frac{1}{2}}|w|^2) }.$$
To prove these formulae for the Szeg\"o kernels, we observe that the
reduced
Szeg\"o kernels are obtained by projecting the Szeg\"o kernel on ${\bf
H}^m$ to
${\bf H}_{\rm red}^m$ as an automorphic kernel, i.e. $$\Pi^\H(x, y) =
\sum_{n
\in {\mathbb Z} } \Pi(x, y\cdot (0, 2\pi n)).$$ Let us write $x = (z, \theta), y
= (w,
\phi)$. Then the $N$-th Fourier component $\Pi^\H_N(x, y)$ of $\Pi^\H$,
i.e.
the projection onto square integrable holomorphic sections of $L^N$, is
given
by: \begin{eqnarray*}\Pi^\H_N(x, y)&=& \int_{{\mathbb R}} e^{- i N t} \Pi( e^{i
t} x, y
) dt\ =\ \int_{{\mathbb R}} e^{- i N t} K(e^{ i t} y^{-1} x ) dt\\&=& \int_{{\mathbb R}}
e^{- i
N t} K(z - w, e^{i (\theta - \phi + t + \Im (z \cdot \bar{w})} )
dt\,.\end{eqnarray*} Here we abbreviated the element $(0, e^{it})$ by
$e^{it}$. Change variables $t \mapsto t- \theta + \phi - \Im (z \cdot
\bar{w})$ to get $$\begin{array}{l}\Pi^\H_N(x, y) = e^{i N (\theta -
\phi)}
e^{i N \Im (z \cdot \bar{w}}) \int_{{\mathbb R}} e^{- i N t} K(z - w, t) dt \\ \\
=
e^{i N ( \theta - \phi)}e^{i N \Im (z \cdot \bar{w})} \hat{K}_t(z - w,
N)\end{array} $$ where $\hat{K}_t$ is the Fourier transform of $K$ with
respect to the $t$ variable. By \cite[p.~585]{Stein}, the full ${\mathbb R}^{2m}
\times {\mathbb R}$ Fourier transform of $K$ is given by $\hat{K} (z, N) = C'_m
e^{-|z|^2/ 2N}$, so by taking the inverse Fourier transform in the $z$
variable we get the Fourier transform just in the $t$ variable:
\begin{equation}\label{PiHN} \Pi^\H_N(x, y) =
\frac{1}{\pi^m} N^m e^{i N (\theta -
\phi)}e^{ iN \Im (z \cdot \bar{w})} e^{- {\frac{1}{2}} N |z -
w|^2}. \end{equation}
(Our constant factor $\frac{1}{\pi^m}$ in (\ref{PiHN}) is determined by
the
condition that $\Pi_N^\H$ is an orthogonal projection.)
In our study of the correlation functions, we will need explicit
formulae for
the horizontal derivatives of the Szeg\"o kernel. The left-invariant
derivatives are given by
\begin{eqnarray} N^{-m} Z^L_q \Pi_N^\H(z,\theta;w,\phi)
&=&\, N (\bar{w}_q - \bar{z}_q) \Pi_N^\H(z,\theta;w,\phi)\,, \nonumber\\
\label{dPiNHleft} & & \\
N^{-m} Z^L_q \bar{W}^L_{ q'} \Pi_N^\H(z,\theta;w,\phi)
&=& N^2 (z_{q'} - w_{q'})(\bar w_q - \bar{z}_q)
\Pi_N^\H(z,\theta;w,\phi) + N\delta_{qq'}
\Pi_N^\H(z,\theta;w,\phi)\,. \nonumber
\end{eqnarray}
Comparing the definitions of the horizontal vector fields with
(\ref{horizontal}), using $h_\H=e^{-|z|^2}$, we see that $d^H_{z_q}=Z^L_q$, as
expected, since $\alpha^L$ agrees with the contact form $\alpha$ for $L_\H$ (as
defined in \S \ref{Szegokernels}).
We
will see later that our formulas for computing correlations are valid
with any
connection, and thus it is sometimes useful to also consider the right
invariant derivatives:
\begin{eqnarray} N^{-m} Z^R_q \Pi_N^\H(z,\theta;w,\phi)
&=&\, N \bar{w}_q \Pi_N^\H(z,\theta;w,\phi)\,, \nonumber\\
\label{dPiNH} & & \\
N^{-m} Z^R_q \bar{W}^R_{ q'} \Pi_N^\H(z,\theta;w,\phi)
&=& N^2 z_{q'}\bar w_q \Pi_N^\H(z,\theta;w,\phi) + N\delta_{qq'}
\Pi_N^\H(z,\theta;w,\phi)\,. \nonumber
\end{eqnarray}
\begin{rem} Recall that the
metric on $\mathcal{O}(N)\to\C\PP^m$ is given by $h^N(z)=(1+|z|^2)^{-N}$ using
the
coordinates and local frame from Example \ref{example1}. Since
$$h^N(z/\sqrt{N})\to h_\H(z)\,,$$ the Heisenberg bundle can be regarded as
the
scaling limit of $\mathcal{O}(N)$. (Of course, in the same way $L_\H$ is the
scaling limit of
$L^N$, for any positive line bundle $L\to M$.)
\end{rem}
\bigskip
\section{Correlation functions} \label{corfuns}
This section begins with a generalization to arbitrary dimension and
codimension a formula of \cite{H} and \cite{BD} for the ``correlation
density function'' in the one-dimensional case. In fact, our formula
(Theorem~\ref{density}) applies to a general class of probability spaces
of
$k$-tuples of (real or complex) functions.
We then specialize to the case where the space of sections
has a Gaussian measure. Finally, we show how the correlations of the
zeros of
$k$-tuples of sections of the $N$-th power of a holomorphic line bundle
are
given by a rational function in the Szeg\"o kernel $\Pi_N$ and its
derivatives (Theorem \ref{npointcor}).
\subsection {General formula for zero correlations}
For our general setting, we let
$(V,h)$ be a Hermitian holomorphic vector bundle on an
$m$-dimensional Hermitian complex manifold $(M,g)$. (Here, we make no
curvature assumptions.) Suppose that
$\mathcal{S}$ is a finite dimensional subspace of the space $H^0(M,V)$ of
global
holomorphic sections of $V$, and let $d\mu$ be a probability measure on
$\mathcal{S}$ given by a semi-positive $\mathcal{C}^0$ volume form that
is
strictly positive in a neighborhood of $0\in\mathcal{S}$. (We shall later
apply
our results to the case where
$V=L^N\oplus\cdots\oplus L^N$, for a holomorphic line bundle
$L$ over a compact complex manifold $M$, and $\mathcal{S}=H^0(M,V)$ with a
Gaussian measure $d\mu$. Our formulation involving general vector
bundles
allows us to reduce the study of $n$-point correlations to the case
$n=1$,
i.e., to expected densities of zeros.)
As in the introduction, we introduce the punctured product
$$M_n=\{(z^1,\dots,z^n)\in \underbrace{M\times\cdots\times
M}_n: z^p\ne z^q \ \ {\rm for} \ p\ne q\}\,,$$ and we write
$$s(z)=(s(z^1),\dots,s(z^n))\,,\
\nabla s(z)=(\nabla s(z^1),\dots,\nabla s(z^n))\,,\quad
z=(z^1,\dots,z^n)\in
M_n\,,$$ where $\nabla s(\zeta)\in T^*_\zeta\otimes V_\zeta$ is the
covariant derivative with respect to the Hermitian connection on $V$. We
define the map
$$\mathcal{J}:M_n \times\mathcal{S}\to \big[({\mathbb C}\oplus T^*_M)\otimes V\big]^n\,,\quad
\mathcal{J} (z,s)=(s(z),\nabla s(z))\,;$$ i.e., $\mathcal{J} (z,s)$ is the 1-jet of
$s$
at $z\in M_n$.
We write $g=\Re\sum g_{qq'}dz_q\otimes d\bar
z_{q'},h_{jj'}=h(e_j,e_{j,})$,
where $\{z_1,\dots,z_m \}$ are local coordinates in $M$
and $\{e_1,\dots,e_k \}$ is a local frame in
$V$ ($m=\dim M,\ k={\rm rank}\,V$). We let $G=\det (g_{qq'})$, $H=\det
(h_{jj'})$. We let
$$d\zeta=\frac{1}{m!}\omega^m_{\zeta}=G(\zeta)\prod_{j=1}^m d\Re \zeta_j
d\Im
\zeta_j\,,\quad \zeta\in M$$ denote Riemannian volume in $M$, and we
write
\begin{equation}\label{dbda} x^p=\sum_j b^p_j e_j(z^p),\quad dx^p=H(z^p)
\prod_{j} d\Re b^p_j d\Im b^p_j \quad x^p\in V_{z^p}\,,\end{equation}
$$\xi^p=\sum_{j,q} a^p_{jq} dz_q\otimes e_j|_{z^p}, \quad d\xi^p=
G(z^p)^{-k}H(z^p)^{m} \prod_{j,q} d\Re a^p_{jq} d\Im
a^p_{jq}
\quad \xi_j\in (T^*_M\otimes V)_{z^p}\,.$$ The quantities $dx^p,\;
d\xi^p$
are the intrinsic volume measures on $V_{z^p}$ and $(T^*_M\otimes
V)_{z^p}$,
respectively, induced by the metrics $g,h$.
\begin{defin} Suppose that $\mathcal{J} $ is surjective. We define the {\it
$n$-point
density\/} $D_n(x,\xi,z)dxd\xi
dz$ of $\mu$ by
\begin{equation}\label{jointdist} \mathcal{J} _*(dz\times
d\mu)=D_n(x,\xi,z)dxd\xi
dz\,,\quad x=(x^1,\dots,x^n)\in
V_{z^1}\times\dots\times V_{z^n}\,,\end{equation}
$$\xi=(\xi^1,\dots,\xi^n)\in
(T^*_M\otimes V)_{z^1}\times\dots\times(T^*_M\otimes V)_{z^n}\,,\
z=(z^1,\dots,z^n)\in M_n\,,$$
$$dx=dx^1\cdots dx^n\,,\quad d\xi=d\xi^1\cdots d\xi^n\,,\quad
dz=dz^1\cdots
dz^n\,.$$ In this case, for each
$z\in M_n$, the (vector-valued) random variable
$(s(z),\nabla s(z))$ has (joint) probability distribution
$D_n(x,\xi,z)dxd\xi$.\end{defin}
\begin{rem} If we let $n=1$ and fix a point $z\in M$, then the measure
$D(x,
\xi, z) dx d\xi$ is intrinsically defined as a measure on the space
$J^1_z(M,
V)$ of 1-jets of sections of $V$ at $z$. Taking a section to its
$1$-jet at
$z$ defines a map $\mathcal{J}_z: \mathcal{S} \to J^1_z(M,V)$ and hence
induces a
measure $\mathcal{J}_{z*} \mu$ on $J^1_z(M,V)$ independently of any choices of
connections, coordinates or metrics. Similarly for $n>1$, $D(x,
\xi, z) dx d\xi$ is an intrinsic measure on $\prod _{p=1}^n J^1_{z_p}(M,
V)$. \end{rem}
For a vector-valued 1-form $\xi\in T^*_{M,z}\otimes V_{z}= {\rm
Hom}(T_{M,z},V_z)$, we let $\xi^*\in {\rm Hom}(V_{z},T_{M,z})$ denote
the adjoint to $\xi$ (i.e., $\langle\xi^*v,t \rangle=\langle v,\xi
t\rangle\,$), and we consider the
endomorphism
$\xi\xi^*\in {\rm Hom}(V_{z},V_{z})$.
In terms of local frames, if $$\xi=\sum_j \xi_j \otimes
e_j=\sum_{j,q}a_{jq}
dz_j\otimes e_j\,,$$ then
$$\xi^*=\sum_{j,q}\alpha_{jq}\frac{\d}{\d z_q}\otimes e_j^*\,,\qquad
\alpha_{jq}=\sum_{j',q'} h_{jj'}\gamma_{q'q}\bar a_{j'q'}\,,$$
where $\big(\gamma_{qq'}\big)=\big(g_{qq'}\big)^{-1}$; hence we have
\begin{equation}\label{endo} \xi\xi^*=\sum_{j,j',j'',q,q'}
h_{j'j''}a_{jq}\gamma_{q'q}\bar a_{j''q'}\, e_j\otimes e_{j'}^*
\,.\end{equation} Its determinant is given by
\begin{equation}\label{detendo}\det(\xi\xi^*)=H\det\left(\sum_{q,q'}
a_{jq}\gamma_{q'q}\bar a_{j'q'}\right)_{1\le j,j'\le k}=H\det \langle\xi_j,
\xi_{j'}\rangle=
H\|\xi_1\wedge\dots\wedge \xi_k\|^2\,.\end{equation}
\begin{rem} The measure $\det(\xi \xi^*) D(0, \xi, z) d\xi dz$ will play
a
fundamental role in our study of correlation functions. We observe here
that it depends only on the metric $\omega$ on $M$, and in the case where
the
zero sets are points ($k=m$), it is
independent of the choice of metric on $M$ as well.
Indeed, as mentioned in the previous remark,
$D(x,\xi, z) dx d\xi$ is well-defined on $J^1_z(M,V)$.
The conditional density $D(0, \xi, z) d\xi$ equals $\mathcal{J}_{z*} \mu /dx
|_{x =
0}$ and thus depends only on the choice of volume forms $dx^p$ on
$V_{z^p}$.
Since $dz/dx$ transforms in the opposite way to $\det \xi \xi^*$ it
follows
that $\det(\xi \xi^*) D(0, \xi, z) d\xi dz$ is an invariantly defined
measure on $(T^*_M\otimes V)^n$. \end{rem}
Recall that for $s\in\mathcal{S}$ so that ${\operatorname{codim\,}} Z_s=k$, we let $|Z_s|$
denote
Riemannian
$(2m-2k)$-volume along the regular points of $Z_s$, regarded as a
measure on
$M$.
\begin{defin} For $s\in\mathcal{S}$ so that ${\operatorname{codim\,}} Z_s=k$, we consider
the product measure on $M_n$,
$$|Z_s|^n=\big(\underbrace{|Z_s|\times\cdots\times|Z_s|}_{n}\big)\,.$$
Its expectation ${\mathbf E}\, |Z_s|^n$ is called the {\it $n$-point zero
correlation measure.} \end{defin}
We shall use the following general formula to compute the correlations
of
zeros and to show universality of the scaling limit:
\begin{theo}\label{density} Let $M,V,\mathcal{S},d\mu$ be as above, and
suppose
that $\mathcal{J} $ is surjective and the volumes $|Z_s|$ are locally
uniformly
bounded above. Then
\begin{equation}\label{a8} {\mathbf E}\,|Z_s|^n =K_n(z)dz \,,\quad
K_n(z)=
\int d\xi\,D_n(0,\xi;z)\prod_{p=1}^n
\det \left(\xi^{p}\xi^{p*}\right)\,.\end{equation}
\end{theo}
The function $K_{n}(z^1,\dots,z^n)$, which is continuous on $M_n$
is called the {\it $n$-point zero correlation function\/}. For $k< m$,
(\ref{a8}) holds on all of the
$n$-fold product $M\times\cdots\times M$, including the diagonal, and
$K_n$ is
locally integrable on $M\times\cdots\times M$ (and is infinite on the
diagonal). In the case
$k=m$, when the zero sets are discrete, the zero correlation measure on
$M\times\cdots\times M$ is the sum of the absolutely continuous measure
$K_n(z)dz$ plus a measure supported on the diagonal.
\medskip\noindent{\em Proof of Theorem \ref{density}:\/}
Consider the Hermitian vector bundle $V_n=\bigoplus_{p=1}^n
\pi_p^*V\longrightarrow M_n$, where $\pi_p:M_n\to M$ denotes the
projection
onto the $p$-th factor. By replacing $V\to M$ with $V_n\to M_n$ and
$s\in H^0(M,V)$ with $$\tilde s(z^1,\dots,z^n)=(s(z^1),\dots,s(z^n))\in
H^0(M_n,V_n)\,,$$ and noting that $T_{M_n,z}=\prod_p T_{M,z^p}$ and
$|Z_s|^n=|Z_{\tilde s}|$, we can assume without loss of generality that
$n=1$.
It follows from the above remarks that $D(0,\xi;z)$ does not depend on
the
choice of connection on
$V$. We can also verify this in terms of local coordinates: write
$s=\sum
b_j e_j$,
$\nabla s =
\sum a_{jq} dz_q \otimes e_j$ as in (\ref{jointdist});
we have $a_{jq}=\frac{\d b_j}{\d z_q}
+ \sum_k b_k\theta^k_{jq}$. Then if we write $a^0_{jq}=\frac{\d b_j}{\d
z_q}$, we have $$\frac{\d(a_{jq},b_j)}{\d(a^0_{jq},b_j)} = 1\,.$$ Hence
$D(0,\xi;z)$ is unchanged if we substitute the (local) flat connection
given
by $a^0_{jq}$.
We now restrict to a coordinate neighborhood $U\subset M$
where $V$ has a local frame $\{e_j\}$. By
hypothesis, we can suppose that the $e_j$ are
restrictions of sections in $\mathcal{S}$.
We write
$s=\sum s_j e_j$, and by the above we may assume that $\nabla s =
\sum ds_j \otimes e_j$.
We use the notation
$$|\!|\!|\xi|\!|\!|=\sqrt{\det(\xi\xi^*)}\,,
\quad {\rm for} \ \xi\in T^*_{M,z}\otimes V_{z}= {\rm
Hom}(T_{M,z},V_z)\,.$$
Then by (\ref{detendo}), $$|\!|\!|\nabla s|\!|\!|^2
=H\|
d s_1\wedge\cdots \wedge d s_k\|^2=\|\Psi\|\,,$$
where $\Psi$ is the $(k,k)$-form on
$U$ given by: $$\Psi=H
\left(\frac{i}{2}
ds_1\wedge
\overline{d s_1}\right)\wedge\cdots\wedge
\left(\frac{i}{2} ds_k\wedge
\overline{d s_k}\right)\,.$$ Thus, by the Leray formula,
\begin{equation}\label{Leray}
|Z_s|=\|ds_1\wedge\cdots\wedge ds_k\|^2 \left.\frac{dz }{
\frac{i}{2}ds_1\wedge d\bar s_1\cdots\wedge\frac{i}{2}ds_k\wedge d\bar
s_k}\right|_{Z_s}=|\!|\!|\nabla s|\!|\!|^2 \left.
\frac{dz}{\Psi}\right|_{Z_s}\,,\end{equation}
Define the measure $\lambda$ on $M\times\mathcal{S}$ by
\begin{equation}\label{lambda} (\lambda,\phi)=\int_\mathcal{S}
\left(|Z_s|,\phi(z,s)\right)d\mu(s)\,.\end{equation} Then
$$\pi_*\lambda={\mathbf E}\,|Z_s|^n\,,$$ where $\pi:M\times\mathcal{S}\to M$ is the
projection.
Hence,
\begin{equation}\label{Leray2}\lambda=\int_\mathcal{S} d\mu(s)\; |Z_s|
=\int_\mathcal{S} d\mu(s)\left.\left(|\!|\!| \nabla s|\!|\!|^2\frac{dz}{
\Psi}\right)\right|_{Z_s}\,.
\end{equation}
For (almost all) $x\in{\mathbb C}^{k}$, let $I(s,x)$ be the measure on $U$ given
by
$$(I(s,x),\phi)=\int_{s(z)=\sum x_j e_j(z)}\phi(z) d{\operatorname{Vol}}_{(2m-2k)n}(z)=
\int_{s(z)=\sum x_j e_j(z)}
|\!|\!| \nabla s|\!|\!|^2\frac{dz}{ \Psi}\phi(z)\,,\ \phi\in\mathcal{C}^0(U) \,,$$
where the second equality is by (\ref{Leray}) applied to $s-\sum x_j
e_j(z)$.
Then
\begin{equation}\label{I} \int
I(s,x) dx=|\!|\!| \nabla s(z)|\!|\!|^2 dz\,.\end{equation}
Now let $\lambda_x$ be the measure on $U$ given by
$$(\lambda_x,\phi)=\int_\mathcal{S} (I(s,x),\phi)d\mu(s) \,.$$
\begin{claim} The map
$x\mapsto (\lambda_x,\phi)$ is continuous.\end{claim}
\noindent To prove this claim, we first note that the hypothesis that
$|Z_s|$
is locally uniformly bounded implies that $(I(s,x),\phi)\le C<+\infty$
uniformly in $s,x$. Thus we can assume without loss of generality that
$\mu$ has compact support in $\mathcal{S}$. By hypothesis, the map
$$\sigma:U\times
\mathcal{S}\to
{\mathbb C}^k,\qquad \sigma(z,s)= (s_1(z),\dots,s_k(z))$$ is a submersion. We can
now write $\lambda_x$ as a fiber integral of a compactly supported $\mathcal{C}^0$
form:
$$\lambda_x=\frac{1}{(m-k)!}\int_{\sigma^{-1}(x)}\phi(z)\omega^{m-k}(z)\wedge
d\mu(s)\,,$$ and thus $\lambda_x$ is continuous, verifying the claim.
\medskip We note that
$\lambda_0=\lambda|_U$. Hence, to complete the proof, we must show that
$$\pi_* \lambda_0=K_1(z)dz|_U\,.$$
By (\ref{jointdist}) and (\ref{I}), for a test function $\phi(x,\xi,z)$,
\begin{eqnarray*}
\int
\phi(x,\xi,z)|\!|\!|\xi|\!|\!|^2 D_1(x,\xi,z)dxd\xi dz &=& \int
d\mu(s)\int \phi(\mathcal{J} (z,s))
|\!|\!| \nabla s(z)|\!|\!|^2dz\\ &=&\int dx\int
(I(s,x),{\phi\circ \mathcal{J} }) d\mu(s)\\ &=&\int
(\lambda_x,\phi\circ \mathcal{J} ) dx\,.\end{eqnarray*}
By choosing $\phi(x,\xi,z)=\rho_\varepsilon(x)\psi(z)$, where $\rho_\varepsilon$ is an
approximate identity, and letting $\varepsilon\to 0$, we conclude that $$\int
\psi(z)K_1(z)dz=\int
\psi(z)|\!|\!|\xi|\!|\!|^2 D_1(0,\xi,z)d\xi dz
=(\lambda_0,\psi(z))\,.$$
\qed\medskip
We note the following analogous formula for real
manifolds:
\begin{theo}\label{density-real} Let $V$ be a $\mathcal{C}^\infty$ real vector
bundle over a
$\mathcal{C}^\infty$ Riemannian manifold $M$, and let $\mu$ be
a probability measure on a finite dimensional vector space
$\mathcal{S}$ of $\mathcal{C}^\infty$ sections of $V$ given by a semi-positive
volume
form that is strictly positive at $0$. Suppose that the volumes $|Z_s|$
are
locally uniformly bounded above. Let $D_n(x,\xi,z)dxd\xi dz$ denote the
$n$-point density of $\mu$. Then
\begin{equation}\label{a8-real} {\mathbf E}\,|Z_s|^n =K_n(z)dz \,,\quad
K_n(z)=
\int d\xi\,D_k(0,\xi,z)\prod_{p=1}^n
\sqrt{\det(\xi^{p}\xi^{p*})}\,.
\end{equation} \end{theo}
The proof is similar to that of Theorem \ref{density}, except that
(\ref{Leray}) is replaced by the Leray formula
\begin{equation}\label{Leray-real} |Z_s|=\|ds_1\wedge\cdots\wedge ds_k\|
\left.\frac{d\zeta }{ ds_1\wedge\cdots\wedge
ds_k}\right|_{Z_s}\end{equation} in the real case.
\subsection{Formula for Gaussian densities}\label{formula-gaussian} We
now
specialize our formula from Theorem
\ref{density} to the case where $\mu$ is a Gaussian measure. Fix
$z=(z^1,\dots,z^n)\in M_n$ and choose local coordinates $\{z^p_q\}$ and
local frames $\{e^p_j\}$ near $z^p$, $p=1,\dots,n$. We consider the
random
variables
$b^p_j,\ a^p_{jq}$ given by
\begin{equation}\label{fg1} s(z^p)=\sum_{j=1}^k b^p_j e^p_j,\quad
\nabla s(z^p)=\sum_{j=1}^k\sum_{q=1}^m a^p_{jq}dz^p_q\otimes
e^p_j,\qquad
p=1,\dots,n.\end{equation} By
(\ref{gaussian})--(\ref{gaussian2}) and (\ref{dbda})--(\ref{jointdist})
the
$n$-point density
$$D_n(x,\xi,z)dxd\xi
dz=D_n\left[\prod_{p=1}^nG(z^p)^{-k}H(z^p)^{m}\right]
dbdadz$$ is given by:
\begin{equation}\label{fg2}D_{n}(b,a;z)=
\frac{\exp\langle-\Delta_{n}^{-1}v,
v\rangle}{\pi^{kn(1+m)}\det\Delta_{n}}\;,\qquad v=\pmatrix b\\
a\endpmatrix\,,
\end{equation}
where \begin{equation}\label{fg3}
\Delta_{n}=\left(
\begin{array}{cc}
A_{n} & B_{n} \\
B^*_{n} & C_{n}
\end{array}\right)\end{equation}
$$A_{n}=\big(A^{jp}_{j'p'}\big)=
\big (\,{\mathbf E}\, b^p_j \bar b^{p'}_{j'}\,\big),\quad
B_{n}=\big(B^{jp}_{j'p'q'}\big)=
\big (\,{\mathbf E}\, b^p_j \bar a^{p'}_{j'q'}\,\big),\quad
C_{n}=\big(C^{jpq}_{j'p'q'}\big)=
\big(\,{\mathbf E}\, a^{p}_{jq}\bar a^{p'}_{j'q'}\,\big);$$
$$j,j'=1,\dots,k;\quad p,p'=1,\dots,n; \quad q,q'=1,\dots,m.
$$
(We note that $A_n,\ B_n,\ C_n$ are $kn\times kn,\ kn\times knm,\
knm\times knm$ matrices, respectively; $j,p,q$ index the rows, and
$j',p',q'$ index the columns.)
The function $D_{n}(0,a;z)$ is a Gaussian function, but it is not
normalized as a probability density. It can be represented as
\begin{equation}\label{fg4}
D_{n}(0,a;z)=Z_{n}(z) D_{\Lambda_n}(a;z),
\end{equation}
where
\begin{equation}\label{fg5}
D_{\Lambda_n}(a;z)=\frac{1}{\pi^{knm}\det\Lambda_{n}}\exp\left(
-{\langle\Lambda^{-1}_{n}a, a\rangle}\right)
\end{equation}
is the Gaussian density with covariance matrix
\begin{equation}\label{fg6}
\Lambda_{n}=C_{n}-B^*_{n}A^{-1}_{n}B_{n} =\left(C^{jpq}_{j'p'q'}
-\sum_{j_1,p_1,j_2,p_2}\bar
B_{jpq}^{j_1p_1}\Gamma^{j_1p_1}_{j_2p_2}B^{j_2p_2}_{j'p'q'}\right)
\qquad (\Gamma=A_n^{-1})
\end{equation}
and
\begin{equation}\label{fg7}
Z_{n}(z)=\frac{\det\Lambda_{n}}{\pi^{kn}\det\Delta_{n}}
=\frac{1}{\pi^{kn}\det A_{n}}\,.
\end{equation}
This reduces formula (\ref{a8}) to
\begin{equation}\label{fg8}
K_{n}(z)=\frac{1}{\pi^{kn}\det A_{n}}\left\langle
\prod_{p=1}^n \det\left(a^{p*}\gamma^pa^p\right)\right\rangle_{\Lambda_{n}}
\end{equation}
where $\langle\cdot\rangle_{\Lambda_{n}}$ stands for averaging with
respect to the
Gaussian density $D_{\Lambda_n}(a;z)$, and $(\gamma^p_{qq'})=
(g^p_{qq'})^{-1}$,
$g^p_{qq'}=g_{qq'}(z^p)$.
\subsection{Densities and the Szeg\"o kernel}\label{D-meet-S} We return
to
our positive Hermitian line bundle $(L,h)$ on a compact complex manifold
$M$
with K\"ahler\ form $\omega=\frac{i}{2}\Theta_h$. We now apply formulas
(\ref{fg3})--(\ref{fg8}) to the vector bundle
$$V=\underbrace{L^N\oplus\cdots\oplus L^N}_k$$
and space of sections $$\mathcal{S}=H^0(M,V)=H^0(M,L^N)^k$$ with the Gaussian
measure $\mu=\nu_N\times\cdots\times\nu_N$, where $\nu_N$ is the
standard
Gaussian
measure on $H^0(M,L^N)$ given by (\ref{gaussian}). We denote the
resulting
$n$-point density by $D_{nk}^N$, and we also write $\Delta_n=\Delta_{nk}^N,\
A_n
=A_{nk}^N$, etc.
As above, we fix
$z=(z^1,\dots,z^n)\in M_n$ and choose local coordinates $\{z^p_q\}$ near
$z^p$, $p=1,\dots,n$. We also choose local frames $\{e^p_L\}$ for $L$
near
the points $z^p$ so that
$$\|e^p_L(z^p)\|_h=1\,.$$ For
$s\in\mathcal{S}$, we write
\begin{equation}\label{dms1} s(z^p)=
\left(
\begin{array}{c}
s_1(z^p)\\
\vdots\\
s_k(z^p)
\end{array}
\right)=
\left(
\begin{array}{c}
b^p_1\\
\vdots\\
b^p_k
\end{array}
\right) (e^p_L(z^p))^{\otimes N}
\,,\end{equation}
\begin{equation}\label{dms2} \nabla_N s_j(z^p)
=\sum_{q=1}^m a^p_{jq}dz^p_q\otimes (e^p_L(z^p))^{\otimes
N}\,.\end{equation}
Since the
$s_j$ are independent and have identical distributions, we have
\begin{equation}\label{dms3} A_{nk}^N=\big(A^{jp}_{j'p'}\big)=
\big(\delta_{jj'}{\mathbf E}\,(b^p_1 \bar
b^{p'}_1)\big)\,,\quad B_{nk}^N=\big(B^{jp}_{j'p'q'}\big)=
\big(\delta_{jj'}{\mathbf E}\,(b^p_1 \bar
a^{p'}_{1q'})\big)\,,
\end{equation}
$$C_{nk}^N=\big(C^{jpq}_{j'p'q'}\big)=
\big(\delta_{jj'}{\mathbf E}\,(a^p_{1q} \bar
a^{p'}_{1q'})\big)\,.$$
We write $$s_1=\sum_{\alpha=1}^{d_N}c_\alpha
S^N_\alpha=\left(\sum_{\alpha=1}^{d_N}c_\alpha f^p_\alpha\right)(e^p_L)^{\otimes
N}\,,$$ where
$\{S^N_\alpha\}$ is an orthonormal basis for $H^0(M,L^N)$.
Using the local coordinates $(z^p,\theta)$ in $X$ as described in \S
\ref{cxgeom}, we have by (\ref{dms3}) and (\ref{PiN}) (noting that
$h(z^p)=0$ by the above choice of local frames),
\begin{equation}\label{dms4}A^{jp}_{j'p'}
=\delta_{jj'}\sum_{\alpha,\beta=1}^{d_N}{\mathbf E}\,(c_\alpha \bar
c_\beta)f^p_\alpha(z^p) \overline{f^{p'}_\beta(z^{p'})}=\delta_{jj'}
\sum_{\alpha=1}^{d_N}f^p_\alpha(z^p) \overline{f^{p'}_\alpha(z^{p'})}=\delta_{jj'}
\Pi_N(z^p,0;z^{p'},0)\,.\end{equation}
Similarly, \begin{equation}\label{dms5}
B^{jp}_{j'p'q'}=\delta_{jj'}
\sum_{\alpha=1}^{d_N}f^p_\alpha(z^p)
\overline{f^{p'}_{\alpha;q'}(z^{p'})}=\delta_{jj'}
d^H_{\bar w_{q'}}\Pi_N(z^p,0;z^{p'},0)\,,\end{equation}
\begin{equation}\label{dms6}
C^{jpq}_{j'p'q'}=\delta_{jj'}
\sum_{\alpha=1}^{d_N}f^p_{\alpha;q}(z^p)
\overline{f^{p'}_{\alpha;q'}(z^{p'})}=\delta_{jj'}
d^H_{z_{q}}d^H_{\bar w_{q'}}\Pi_N(z^p,0;z^{p'},0)\,.\end{equation}
\begin{lem}\label{detN} There is a positive integer $N_0=N_0(M,n)$ such
that
$$\det\big(\Pi_N(z^p,0;z^{p'},0)\big)_{1\le p,p'\le n} \ne 0\,,$$
for distinct points $z^1,\dots,z^n$ of $M$ and for all $N\ge N_0$.
\end{lem}
\begin{proof} It is a well-known consequence of the Kodaira Vanishing
Theorem
(see for example,
\cite{GH}) that we can find $N_0$ such that if $N\ge N_0$ and
$x_1,\dots,x_n\in M$ with $x_p\ne x_1$ for $2\le p\le n$, then there is
a
section $s\in H^0(M,L^N)$ with $s(x_1)\ne 0$ and $s(x_p)=0$ for $2\le
p\le
n$.
We write $\widetilde A_{pp'}=\Pi_N(z^p,0;z^{p'},0)$. Suppose on the
contrary that
$\det (\widetilde A_{pp'})=0$, and chose a nonzero vector
$v=(v_1,\dots,v_n)$ such that $\sum_p v_p\widetilde A_{pp'}=0$.
Then recalling (\ref{szego}), we have
\begin{equation}\label{det1} 0=\sum_{p,p'} v_p \widetilde A_{pp'}\bar v_{p'}
= \sum_{p,p',\alpha} v_p \widehat S^N_\alpha(z^p,0) \overline{\widehat
S^{N}_\alpha(z^{p'},0)
v_{p'}} =\sum_{\alpha=1}^{d_N} |x_\alpha|^2\,,\end{equation}
where $x_\alpha=\sum_p v_p\widehat S^N_\alpha(z^p,0)$. Since the $S^N_\alpha$ span
$H^0(M,L^N)$, it follows that for all $s\in H^0(M,L^N)$, we have $\sum_p
v_p \hat s(z^p) =0$.
But this contradicts the fact that, choosing $p_0$ with $v_{p_0}\ne
0$, we can find a section $s\in H^0(M,L^N)$ with
$s(z^{p_0})\ne 0$ and $s(z^p)=0$ for $p\ne p_0$.
\end{proof}
Thus we see that the $n$-point correlation functions depend only on the
Szeg\"o kernel, as follows:
\begin{theo}\label{npointcor} Let $(L,h)$ be a positive Hermitian line
bundle
on an
$m$-dimensional compact complex manifold $M$ with K\"ahler\ form
$\omega=\frac{i}{2}\Theta_h$, let
$\mathcal{S}=H^0(M,L^N)^k$ ($k\ge 1$), and give $\mathcal{S}$ the standard Gaussian
measure
$\mu$ described above. Let $n\ge 1$ and suppose that $N$ is sufficiently
large so that
$\mathcal{J}$ is surjective. Let $z=(z^1,\dots,z^n)\in M_n$ and choose local
coordinates $(\zeta_1,\dots,\zeta_m)$ at each point $z^p$ such that
$\Theta_h(z^p)=\sum_q d\zeta_q\wedge d\bar\zeta_q$, $1\le p\le n$.
Then the $n$-point correlation $K_{nk}^N(z)$ is given by a universal
rational
function, homogeneous of degree $0$, in the values of $\Pi_N$ and its
first
and second derivatives at the points $(z^p,z^{p'})$. Specifically,
\begin{equation}\label{npointcoreq}
K_{nk}^N(z)= \frac{\mathcal{P}_{nkm}\big(\Pi_N(z^p,z^{p'}),d^H_{\bar
w_{q}}\Pi_N(z^p,z^{p'}), d^H_{z_{q}}\Pi_N(z^p,z^{p'}),
d^H_{z_{q}}d^H_{\bar
w_{q'}}\Pi_N(z^p,z^{p'})\big)}{\pi^{kn}\left[\det
\big(\Pi_N(z^p,z^{p'})\big)_{1\le p,p'\le
n}\right]^{k(n+1)}}\end{equation}
($1\le p,p'\le
n,\ 1\le q,q'\le m$), where $\mathcal{P}_{nkm}$ is a universal homogeneous
polynomial of degree $kn(n+1)$ with integer
coefficients depending only on $n,k,m$.
\end{theo}
\begin{proof} The $n$-point zero correlation $K_{nk}^N(z)$ is given by
equation (\ref{fg8}) with
$\gamma^p_{qq'}=\delta_{qq'}$.
By the Wick formula (\cite[(I.13)]{Si}), the expectation
$$\left\langle
\prod_{p=1}^n \det\left(a^{p*}a^p\right)\right\rangle_{\Lambda_{n}}$$ in
(\ref{fg8}) is a homogeneous polynomial (over ${\mathbb Z}$) of degree $kn$ in
the
coefficients of
$\Lambda_{n}$. By (\ref{fg6}) and (\ref{dms4}), the coefficients of $\det
\big(\Pi_N(z^p,z^{p'})\big)\Lambda_n$ are homogeneous polynomials of degree
$n+1$
in the coefficients of $A_n,B_n,C_n$. The conclusion then follows from
(\ref{dms4})--(\ref{dms6}).
\end{proof}
\begin{rem}
In the statement of Theorem \ref{npointcor}, we wrote $\Pi_N(z,w)$ for
$\Pi_N(z,\theta;w,\phi)$. Since the expression is homogeneous of degree
0, it
is independent of $\theta$ and $\phi$. Alternately, we could regard
$\Pi_N(z,w)$ as functions on $M\times M$ having values in
$L_z\otimes\overline{L_w}$ (replacing the horizontal derivatives with
the
corresponding covariant derivatives); again the degree 0 homogeneity
makes the
expression a scalar. Furthermore, since Theorem \ref{density} is valid
for
all connections, we can replace the horizontal derivatives in
(\ref{npointcoreq}) with the derivatives with respect to an arbitrary
connection.\end{rem}
\subsection {Zero correlation for
${\operatorname{SU}}(m+1)$-polynomials} \label{zcorpoly}
In this section, we use our methods to describe the zero correlation
functions for ${\operatorname{SU}}(m+1)$-polynomials. We do not carry out the
computations in
complete detail, since we are primarily interested in the scaling
limits,
which we shall compute in \S \ref{formulas}.
The ${\operatorname{SU}}(m+1)$-polynomials are random homogeneous
polynomials of degree $N>0$ on ${\mathbb C}^{m+1}$,
\begin{equation}\label{a1}
s(z)=s(z_0,z_1,\dots,z_m)=\sum_{|J|=N}
\sqrt{N!/J!}\,{c_J}z^J,\quad z^J=z_0^{j_0}\cdots
z_m^{j_m},\quad J!=j_0!\cdots j_m!,
\end{equation}
where the coefficients $c_J$ are complex independent Gaussian random
variables with mean 0 and variance 1:
\begin{equation}\label{a2}
{\mathbf E}\, c_J=0;\qquad {\mathbf E}\, c_J\overline{c_K}=
\delta_{JK},\quad \delta_{JK}=\delta_{j_0k_0}\dots\delta_{j_mk_m};
\qquad {\mathbf E}\, c_Jc_K=0.
\end{equation}
Then $s(z)$ is a Gaussian random polynomial on
${\mathbb C}^{N+1}$ with first and second moments given by
\begin{equation}\label{a3}
{\mathbf E}\, s(z)=0;\qquad {\mathbf E}\, s(z)\overline{s(w)}=\langle z, w
\rangle^N =\left(\sum_{q=0}^m z_q\overline{w_q}\right)^N;
\qquad {\mathbf E}\, s(z)s(w)=0.
\end{equation}
This implies that
the probability distribution of $s(z)$ is invariant with respect
to the map $s(z)\to s(Uz)$ for all $U\in {\operatorname{SU}}(m+1)$.
Let $(\mathcal{S}_N,\mu_N)$ denote the Gaussian probability space of
independent $k$-tuples ($k\le m$) of
${\operatorname{SU}}(m+1)$-polynomials of degree $N$. For
$s=(s_1,\dots,s_k)\in \mathcal{S}_N$, the zero set
\[
Z_s=\{z: s_1(z)=\dots=s_k(z)=0\}\,.
\]
is an algebraic variety in the complex projective space $\C\PP^m$.
We will assume that $\C\PP^m$ is supplied with the Fubini-Study
Hermitian metric $\omega$, which is ${\operatorname{SU}}(m+1)$-invariant. In the affine
coordinates $z=(1,z_1,\dots,z_m)$,
\begin{equation}\label{a3a}
\omega=\frac{\sqrt{-1}}{ 2}\partial\dbar\log\left(1+\sum|z_q|^2\right)=\frac
{\sqrt{-1}}{
2}\left[\frac{\sum dz_q\wedge\overline{dz_q}}{1+\sum |z_q|^2}
-\frac{\left(\sum \overline{z_q}dz_q\right)\wedge \left(\sum
z_q\overline{dz_q}\right)} {\left(1+\sum |z_q|^2\right)^2}\right]\,;
\end{equation}
i.e.,
\begin{equation}\label{FSg} \omega=
\frac{\sqrt{-1}}{ 2}\sum g_{qq'}dz_q\wedge
dz_{q'}\,, \quad
g_{qq'}=\frac{(1+|z|^2)\delta_{qq'}-\bar
z_qz_{q'}}{(1+|z|^2)^2}\,.\end{equation}
To simplify our computations, we consider only points $z^p$ with finite
affine
coordinates, $z^p=(1,z^p_1,\dots,z^p_m),\; p=1,\dots,n$, and we regard
the
${\operatorname{SU}}(m+1)$-polynomials $s_j$ as polynomials of degree $\le N$ on ${\mathbb C}^m$;
i.e.,
we regard the $s_j$ as sections of the trivial line bundle on ${\mathbb C}^m$
with
the flat metric $h=1$ (so that the covariant derivatives coincide with
the
usual derivatives of functions).
As above, we consider the
random variables $$b^p_j=s_j(z^p),\; a^p_{jq}=\frac{\d s_j }{ \d
z_q}(z^p)\,,$$
and we denote their joint
distribution by
\begin{equation}\label{a11a}
\begin{array}{rl}
D_{nk}^N(b,a;z)db\,da,\quad & b=\left(b^1,\dots,b^n\right),\quad
b^p=
(b^p_j)_{j=1,\dots,k};\\
&\\
& a=\left(a^1,\dots,a^n\right),\quad
a^p=\left(a^p_{jq}\right)_{j=1,\dots,k;\;q=1,\dots,m}\,.
\end{array}
\end{equation}
(Here, the $n$-point density is with respect to Lebesgue measure
$db=\prod
db^p_j d\bar b^p_j,\; da=\prod da^p_{jq}d\bar a^p_{jq}$.) We assume that
$N>nm$ to ensure that $\mu_N$
possesses a continuous $n$-point density.
Since $\mu_N$ is Gaussian, the density $D_{nk}^N(b,a;z)$ is Gaussian as
well,
and it is described by the covariance matrix
\begin{equation}\label{a14}
\Delta_{nk}^N=\left(
\begin{array}{cc}
A_{nk}^N & B_{nk}^N \\
B_{nk}^{N*} & C_{nk}^N
\end{array}\right)
\end{equation}
where
\begin{equation}\label{a14a}
\begin{array}{l}
A_{nk}^N=\left (\,{\mathbf E}\, s_j(z^p)\overline{s_{j'}(z^{p'})}\,\right),\\
B_{nk}^N=\left (\,{\mathbf E}\, s_j(z^p)\overline{\frac{\d s_{j'}}{\d
z_{q'}}(z^{p'})}\,\right),\\
C_{nk}^N=\left (\,{\mathbf E}\, \frac{\d s_j}{\d z_q}(z^p)
\overline{\frac{\d s_{j'}}{\d
z_{q'}}(z^{p'} )}\,\right);\\
j,j'=1,\dots,k;\quad p,p'=1,\dots,n; \quad q,q'=1,\dots,m.
\end{array}
\end{equation}
By (\ref{a14a}) and (\ref{a3}),
\begin{equation}\label{a23}
\begin{array}{ll}
A_{nk}^N=\left(\delta_{jj'}S_N(z^p,z^{p'})\right)\,,&\displaystyle
S_N(z,w)=\left(1+\sum_{r=1}^m z_r\overline{w_r}\right)^N,\\
B_{nk}^N=\left(\delta_{jj'}S_{Nq'}(z^p,z^{p'})\right)\,,
&\displaystyle
S_{Nq'}(z,w)=Nz_{q'}\left(1+\sum_{r=1}^m
z_r\overline{w_r}\right)^{N-1}\,, \\
C_{nk}^N=\left(\delta_{jj'} S_{Nqq'}(z^p,z^{p'})\right)\,,&\displaystyle
S_{Nqq'}(z,w)=N(N-1)\overline{w_q}z_{q'}
\left(1+\sum_{r=1}^m z_r\overline{w_r}\right)^{N-2}\\
&\displaystyle \qquad +\delta_{qq'}N\left(1+\sum_{r=1}^m
z_r\overline{w_r}\right)^{N-1}
\end{array}
\end{equation}
The $n$-point zero correlation functions $K_{nk}^N$ for the
${\operatorname{SU}}(m+1)$-polynomial $k$-tuples $\mathcal{S}_{N}$ can be computed by
substituting
(\ref{a23}) into formulas (\ref{fg6}) and (\ref{fg8}). (Alternately, we
can
compute the zero correlation functions with respect to the Euclidean
volume
on ${\mathbb C}^m$ by setting $\gamma^p=$Id in (\ref{fg8}).)
\begin{rem} Note that the one-point correlation
function, or the zero-density function, is constant, since it is
invariant
with respect to the group ${\operatorname{SU}}(m+1)$. Indeed,
by B\'ezout's theorem and
(\ref{volmeasure}),
\begin{equation}\label{a4}
|Z_s|(1)={\operatorname{Vol}}(V_s)=\int_{V_s}{\textstyle\frac{1}{ (m-k)!}}\omega^{m-k}
=\Omega_{2m-2k}\deg V_s=\Omega_{2m-2k}N^k\;,
\end{equation}
where \begin{equation}\label{a4a}\Omega_{2\ell} = {\operatorname{Vol}}\, \C\PP^\ell =
\frac{\pi^\ell}{
\ell !}\,.\end{equation} Hence,
\begin{equation}\label{a32b}
K_{1k}^N(z)=\frac{{\operatorname{Vol}}\, Z_s}{{\operatorname{Vol}}\,
\C\PP^m}=\frac{N^{k}\Omega_{2m-2k}}{\Omega_{2m}}
=\frac{N^{k}m!}{(m-k)!\pi^k}\,.\end{equation} We can also use our
formulas to
compute $K_{1k}^N$ directly: By (\ref{a23}),
\begin{equation}\label{a32c}
A_{1k}^N=\Big(\delta_{jj'}(1+|z|^2)^N\Big)\,,\quad
B_{1k}^N=\Big(\delta_{jj'}Nz_{q'}(1+|z|^2)^{N-1}\Big)\,,\end{equation}
$$C_{1k}^N=\Big(\delta_{jj'}N[(N-1)\bar
z_qz_{q'}+(1+|z|^2)\delta_{qq'}](1+|z|^2)^{N-2}\Big)\,.$$ Hence by
(\ref{fg6}),
\begin{equation}\label{a32d} \Lambda_{1k}^N =
\Big(\delta_{jj'}N[(1+|z|^2)\delta_{qq'}-\bar
z_qz_{q'}](1+|z|^2)^{N-2}\Big)=
\Big(\delta_{jj'}N(1+|z|^2)^Ng_{qq'}(z)\Big)\,.\end{equation}
In the hypersurface case ($k=1$), we compute
$$K_{11}^N=\frac{1}{ \pi (1+|z|^2)^N}\left.\left\langle\sum_{q,q'=1}^m
\bar a_{1q}\gamma_{qq'} a_{1q'}\right\rangle\right|_{\Lambda_{11}^N}=
\frac{N}{
\pi}\sum_{q,q'=1}^m \gamma_{qq'}g_{q'q}=\frac{Nm}{ \pi}\,,$$
as expected. For $k>1$, we have
$\Lambda_{1k}^N(0)=NI$ where
$I$ is the unit matrix, and (\ref{fg8}) yields
$$
K_{1k}^N(0)=\frac{N^{k}}{\pi^k}\left\langle
\det\left(\sum_{q=1}^m
\bar a_{jq}{a_{j'q}}\right)_{j,j'=1,\dots,k} \right\rangle_I
=\frac{N^{k}m!}{(m-k)!\pi^k}\,,$$
which agrees with (\ref{a32b}).
\end{rem}
\bigskip
\section{Universality and scaling}
Our goal is to derive scaling limits of the $n$-point correlations
between the
zeros of {\it random} $k$-tuples of sections of powers of a positive
line
bundle over a complex manifold. We expect the scaling limits to exist
and to
be universal in the sense that they should depend only on the dimensions
of
the algebraic variety of zeros and the manifold. Our plan is the
following. We first describe scaling in the Heisenberg model, which we
use
to provide the universal scaling limit for the
Szeg\"o kernel (Theorem \ref{neardiag}). Together with Theorem
\ref{npointcor}, this demonstrates the universality of the scaling-limit
zero
correlation in the case of powers of any positive line bundle on any
complex
manifold.
\subsection{Scaling of the Szeg\"o kernel in the Heisenberg group}
Our model for scaling is the Szeg\"o kernel for the reduced Heisenberg
group
described in
\S \ref{example2}. Recall that for the simply-connected Heisenberg
group
${\bf H}^m$, the scaling operators
(or Heisenberg dilations)
$$ \delta_r (\zeta, t) = (r \zeta, r^2 t)\,,\quad r
\in {\mathbb R}^+$$
are automorphisms of ${\bf H}^m$ (\cite{F} \cite{Stein}). Since the
Szeg\"o kernel $\Pi$ of ${\bf H}^m$ is the unique self-adjoint
holomorphic
reproducing kernel, it follows that it must be invariant (up to a
multiple) under these automorphisms. In fact, one has
(\cite[p.~538]{Stein}):
\begin{equation} \label{SI}
\Pi(\delta_rx,\delta_ry)=r^{-2m-2}\Pi(x,y)\end{equation}
The condition for a dilation $\delta_r$ to descend to the quotient
group ${\bf H}_{\rm red}^m$ is that
$r^2 {\mathbb Z} \subset {\mathbb Z}$, or equivalently,
$r= \sqrt{N}$
with $N \in {\mathbb Z}^+$. Note however that $\delta_{\sqrt{N}}$ is not an
automorphism of ${\bf H}^m_{\rm red}$ and there
is no well-defined dilation by $\sqrt{N}^{-1}$.
The scaling identity (\ref{SI}) descends to ${\bf H}^m_{\rm red}$ in
the form \begin{equation}\Pi^\H_N(x, y) = N^m \Pi^\H_1 (\delta_{\sqrt{N}}\,
x,
\delta_{\sqrt{N}}\, y)\end{equation}
with \begin{equation}\label{sksl} \Pi^\H_1(x, y) =\frac{1}{\pi^m}
e^{i(\theta - \phi)}e^{ i \Im (z \cdot \bar{w})} e^{-
{\frac{1}{2}} |z - w|^2}\,,\quad x=(z,\theta)\,,\ y=(w,\phi)\,. \end{equation}
(Recall (\ref{PiHN}).) Informally, we may say that the scaling limit of
$\Pi^\H_N$
equals $\Pi^\H_1$. Since scaling by $\sqrt{N}^{-1}$ is not well-defined
on ${\bf H}_{\rm red}^m$ it is more correct to say that $\Pi^\H_N$ is
the
$\sqrt{N}$ scaling of the scaling limit kernel.
\subsection{Scaling limit of a general Szeg\"o kernel}
We now show that $\Pi^\H_1$ is the scaling limit of the
$N$-th Szeg\"o kernel $\Pi_N$ of an arbitrary positive line bundle $L
\to
M$ in
the sense of the following ``near-diagonal asymptotic estimate for the
Szeg\"o kernel."
\begin{theo} \label{neardiag}
Let $z_0\in M$ and choose local coordinates in a neighborhood of
$z_0$ so that $\Theta_h(z_0)=\sum dz_j\wedge d\bar z_j$. Then
\begin{eqnarray*}N^{-m}\Pi_N(z_0+\frac{u}{\sqrt{N}},\frac{\theta}{N};
z_0+\frac{v}{\sqrt{N}},\frac{\phi}{N})&=&
\frac{1}{\pi^m} e^{i(\theta-\phi)+i\Im
(u\cdot \bar v)-{\frac{1}{2}} |u-v|^2} + O(N^{-1/2})\\ &=&
\Pi^\H_1(u,\theta;v,\phi) + O(N^{-1/2})\;.\end{eqnarray*}
\end{theo}
To prove Theorem \ref{neardiag}, we need to recall the Boutet de
Monvel-Sjostrand parametrix construction:
\begin{theo} {\rm \cite [Th.~1.5 and \S 2.c]{B.S}} Let $\Pi(x,y)$ be
the
Szeg\"o kernel of the boundary $X$ of a bounded strictly pseudo-convex
domain
$\Omega$ in a complex manifold $L$. Then: there exists a symbol $s \in
S^{n}(X \times X \times {\mathbb R}^+)$ of the type $$s(x,y,t) \sim
\sum_{k=0}^{\infty} t^{n-k} s_k(x,y)$$ so that $$\Pi (x,y) =
\int_0^{\infty}
e^{i t\psi(x,y)} s(x,y,t ) dt $$ where the phase $\psi \in C^{\infty}(X
\times X)$ is determined by the following properties:
$\bullet$ $\psi(x,x) = \frac{1}{i} \rho(x)$ where $\rho$ is the defining
function of $X$.
$\bullet$ $\bar{\partial}_x \psi$ and $\partial_y \psi$ vanish to
infinite
order along the diagonal.
$\bullet$ $\psi(x,y) = -\bar{\psi}(y,x)$. \end{theo}
The integral is defined as a complex oscillatory integral and is
regularized
by taking the principal value (see \cite{B.S}). The phase is determined
only
up to a function which vanishes to infinite order at $x = y$ and its
Taylor
expansion at the diagonal is given by \begin{equation} \psi(x + u, x +
v) =
\frac{1}{i} \sum \frac{\partial^{\alpha + \beta} \rho} {\partial
z^{\alpha}\partial \bar{z}^{\beta}}(x)
\frac{u^{\alpha}}{\alpha!}\frac{\bar{v}^{\beta}}{\beta!}.\end{equation}
The Szeg\"o kernels $\Pi_N$ are Fourier coefficients of $\Pi$ and hence
may be
expressed as: \begin{equation}\Pi_N(x,y) = \int_0^{\infty} \int_0^{2\pi}
e^{- i
N \theta} e^{it \psi( r_{\theta} x,y)} s(r_{\theta} x,y,t)
d\theta dt \end{equation} where $r_{\theta}$ denotes the $S^1$ action
on $X$. Changing variables $t \mapsto N t$ gives
\begin{equation} \Pi_N(x,y) = N \int_0^{\infty} \int_0^{2\pi}
e^{ i N ( -\theta + t \psi( r_{\theta} x,y))} s(r_{\theta} x,y, t
N) d\theta dt\,.\end{equation}
We now fix $z_0$ and consider the asymptotics of
\begin{equation}\begin{array}{l}\displaystyle \Pi_N(z_0 +
\frac{u}{\sqrt{N}}, 0; z_0 + \frac{v}{\sqrt{N}}, 0)
\\[14pt]\displaystyle
\quad\quad=
N \int_0^{\infty} \int_0^{2\pi}
e^{ i N ( -\theta + t\psi( z_0 + \frac{u}{\sqrt{N}}, \theta; z_0 +
\frac{v}{\sqrt{N}}, 0))} s(z_0 + \frac{u}{\sqrt{N}}, \theta; z_0 +
\frac{v}{\sqrt{N}}, 0), t N) d\theta dt \,.\end{array}\end{equation}
In our setting the phase takes the following concrete form: We
let $h (z,\bar{w})$ be the almost analytic function on $M \times M$
satisfying
$h(z,\bar{z}) = \|e_L\|^{-2}_h(z)$. The
function $h(z,\bar{w})$ is defined by \begin{equation}\label{taylorh}
h (z_0 + u, \bar z_0 +
\bar v) = \sum \frac{\partial^{\alpha + \beta} h(z,\bar{z})} {\partial
z^{\alpha} \partial \bar{z}^{\beta}} (z_0) \frac{u^{\alpha}}{\alpha!}
\frac{\bar{v}^{\beta}}{\beta!}. \end{equation}
We consider the complex manifold $Y=L^*$
and we let $(z,\lambda)$ denote the
coordinates of $\xi \in Y$ given by $\xi = \lambda (e_L^*)_z$. In the
associated coordinates $(x,y) = (z, \lambda, w, \mu)$ on $Y \times Y$,
we
have: \begin{equation} \rho(z,\lambda) = 1 - h(z,
\bar{z})|\lambda|^2,\;\;\;\; \psi(z, \lambda, w, \mu) = \frac{1}{i}(1 -
h(z,\bar{w}) \lambda \bar{\mu})\;. \end{equation} We consider
$\Omega=\{\rho<0\}$ and $X=\partial\Omega=\{\rho=0\}$. We may assume
without
loss of generality that $h(z,\bar{w}) = \bar{h}(w,\bar{z})$ since $h(z,
\bar{z})$ is real so we could replace $h$ by $\frac{1}{2}h(z,\bar{w}) +
{\frac{1}{2}}\bar{h}(w,\bar{z})$. On $X$ we have $h(z, \bar{z}) |\lambda|^2 =
1$ so
we may write $\lambda = h(z, \bar{z})^{-{\frac{1}{2}} } e^{i \phi}$, and
similarly
for $\mu$. So for $(x,y) = (z, \phi, w, \phi') \in X \times X$ we have
\begin{equation} \psi(z, \phi, w, \phi') = \frac{1}{i} \left[1 -
\frac{h(z,\bar{w})}{ \sqrt{h(z, \bar{z})} \sqrt{h(w, \bar{w})}} e^{i
(\phi -
\phi')}\right]\;. \end{equation} It follows that
\begin{equation}\begin{array}{l}\displaystyle\psi( z_0 +
\frac{u}{\sqrt{N}},
\theta; z_0 + \frac{v}{\sqrt{N}}, 0)\\[14pt] \displaystyle \quad\quad=
\frac{1}{i} \left[1 - \frac{h(z_0 + \frac{u}{\sqrt{N}},\bar{z_0} +
\frac{\bar{v}}{\sqrt{N}})}{ \sqrt{h(z_0 + \frac{u}{\sqrt{N}}, \bar{z_0}
+
\frac{\bar u}{\sqrt{N}})} \sqrt{h(z_0 + \frac{v}{\sqrt{N}}, \bar{z_0} +
\frac{\bar v}{\sqrt{N}})}} e^{i \theta}\right] .
\end{array}\end{equation}
We now assume that $e_L$ is a normal frame centered at
$z_0$. By definition, this means that \begin{equation} h(z_0)=1,\quad
\d h |_{z_0} = \bar\partial h|_{z_0} = 0. \end{equation} We furthermore assume
that
our coordinates $\{z_j\}$ are chosen so that the Levi form of $h$ is
the identity at $z_0$:
\begin{equation}\frac{\d^2 h}{\d z^\alpha \d z^\beta}(z_0,\bar z_0)
=\delta_{\alpha\beta}\,.\end{equation} (This is equivalent to specifying that
$\omega(z_0)=\frac{i}{2}\sum_j dz_j \wedge d\bar z_j$.) Then by
(\ref{taylorh}),
\begin{equation}\label{taylorh2} h(z_0 + \frac{u}{\sqrt{N}},\bar z_0 +
\frac{\bar v}{\sqrt{N}}) = \frac{1}{N}u\cdot \bar v +
O(N^{-3/2}). \end{equation}
Now let us return to the phase. It is given by
\begin{equation}\label{phase}
t \left[ 1 - \frac{h(z_0 + \frac{u}{\sqrt{N}},\bar z_0 + \frac{\bar
v}{\sqrt{N}})}{h(z_0 + \frac{u}{\sqrt{N}},\bar z_0 + \frac{\bar
u}{\sqrt{N}})^{{\frac{1}{2}}} h(z_0 + \frac{v}{\sqrt{N}},\bar z_0 + \frac{\bar
v}{\sqrt{N}})^{{\frac{1}{2}}}} e^{i \theta}\right] - i \theta. \end{equation} By
(\ref{taylorh2}), the phase (\ref{phase}) has the form: \begin{equation}
(t[ 1
- e^{i \theta}] - i \theta) + \frac{t}{N}\left[u\cdot\bar v - {\frac{1}{2}}
|u|^2 -
{\frac{1}{2}} |v|^2\right]e^{i \theta} + O(N^{-3/2}). \end{equation} It is now
evident
that $\Pi_N(z_0 + \frac{u}{\sqrt{N}}, 0; z_0 + \frac{v}{\sqrt{N}}, 0)$
is
given by an oscillatory integral with phase $(t [ 1 - e^{i \theta}] - i
\theta)$;
the latter two terms can be absorbed into the amplitude.
Thus we have: \begin{equation}\begin{array}{l} \Pi_N(z_0 +
\frac{u}{\sqrt{N}}, 0; z_0 + \frac{v}{\sqrt{N}}, 0)\\[14pt]\quad =
N \int_0^{\infty} \int_0^{2\pi}
e^{ i N (t [ 1 - e^{i \theta}] - i \theta)} e^{t[u\cdot\bar v - {\frac{1}{2}}
|u|^2 -
{\frac{1}{2}} |v|^2] +O(N^{-1/2})} s(z_0 + \frac{u}{\sqrt{N}}, \theta; z_0
+ \frac{v}{\sqrt{N}}, 0), t N) d\theta
dt.\end{array}\end{equation} We may then evaluate the integral
asymptotically by the stationary phase method as in \cite{Z}. The phase
is
precisely the same as occurs in $\Pi_N(x,x)$, and as discussed in
\cite{Z}, the single critical point occurs at $t = 1, \theta = 0$. We
may
also Taylor-expand the amplitude to determine its contribution to the
asymptote.
Precisely as in the calculation of the stationary phase expansion in
\cite{Z},
we get: \begin{equation} \Pi_N(z_0 +
\frac{u}{\sqrt{N}}, 0; z_0 + \frac{v}{\sqrt{N}}, 0) =
\frac{N^m}{\pi^m} e^{u\cdot\bar v - {\frac{1}{2}} |u|^2 -
{\frac{1}{2}} |v|^2} + O(N^{m - {\frac{1}{2}}}). \end{equation}
Finally, we note that $$u\cdot\bar v - {\frac{1}{2}} |u|^2 -
{\frac{1}{2}} |v|^2=-{\frac{1}{2}} |u-v|^2 +i\Im (u\cdot \bar v)\,,$$
which completes the proof
of Theorem \ref{neardiag}.\qed
\subsection{Universality of the scaling limit of correlations of zeros}
We are now ready to pass to the scaling limit as $N\to\infty$ of the
correlation functions of
sections of powers $L^N$ of our line bundle. To
explain
this notion, let us consider the case $k=m$ where the zeros are (almost
surely) discrete. An $m$-tuple of sections of $L^N$ will have $N^m$
times
as many zeros as $m$-tuples of sections of $L$. Hence we must expand
our
neighborhood (or contract our ``yardstick") by a factor of $N^{m/2}$.
Let $z^0\in M$ and choose a coordinate neighborhood $U\in M$ with
coordinates
$\{z_j\}$ for which
$z^0=0$ and
$\omega(z^0)=\frac{i}{2}\sum_q dz_q\wedge d\bar z_q$. We define the
{\it
$n$-point scaling limit zero correlation function\/}
$$K_{nkm}^\infty(z)=\lim_{N\to\infty}\frac{1}{ N^{nk}}
K_{nk}^N\left(\frac{z}{
\sqrt{N}}\right)\,,\quad z=(z^1,\dots,z^n)\in ({\mathbb C}^m)_n\,.$$
We show below (Theorem \ref{uslc}) that this limit exists and that
$K_{nkm}^\infty$ is universal by passing to the limit in
Theorem \ref{npointcor}, using Theorem \ref{neardiag}. First, we need
the
following fact:
\begin{lem}\label{detinfty} Let $z^1,\dots,z^n$ be distinct points
of ${\mathbb C}^m$. Then
$$\det\left(\Pi^\H_1(z^p,0;z^{p'},0)\right)= e^{-\sum|z^p|^2}\det\left(
e^{z^p\cdot\bar z^{p'}}\right) \ne 0\;.$$\end{lem}
\begin{proof} We consider the first Szeg\"o projector on the reduced
Heisenberg group
\begin{equation} \Pi^\H_1:\mathcal{L}^2(\H^m_{\rm red})\to \mathcal{H}^2_1(\H^m_{\rm
red})\approx \mathcal{L}^2({\mathbb C}^m,e^{-|z|^2})\cap\mathcal{O}({\mathbb C}^m)\,,\end{equation}
where
$$
\mathcal{L}^2({\mathbb C}^m,e^{-|z|^2})=\left\{f\in\mathcal{L}^2_{\rm loc}({\mathbb C}^m):
{\textstyle \int_{{\mathbb C}^m}}|f|^2 e^{-|z|^2}dz <+\infty\right\}\,.$$
(See the remark at the end of \S \ref{example2}.) Its kernel can be
written
in the form \begin{equation}
\Pi^\H_1(z,\theta;w,\phi)=e^{i(\theta-\phi)}
\sum_{\alpha=1}^\infty f_\alpha(z)\overline{f_\alpha(w)}\,,\end{equation}
where the $f_\alpha$ form a complete orthonormal basis for
$\mathcal{L}^2({\mathbb C}^m,e^{-|z|^2})\cap\mathcal{O}({\mathbb C}^m)$. (E.g., $\{f_\alpha\}$ can be
taken
to be the set of monomials $\left\{c_{j_1\cdots j_m}z_1^{j_1}\cdots
z_m^{j_m}\right\}$. In fact, $\Pi^\H_1(z,0;w,0))$ is just a ``weighted
Bergman kernel" on ${\mathbb C}^m$.) We now mimic the proof of Lemma \ref{detN},
except this time we have an infinite sum over the index $\alpha$; this sum
converges uniformly on bounded sets in ${\mathbb C}^m\times {\mathbb C}^m$ since the sup
norm
over a bounded set is dominated by the Gaussian-weighted $\mathcal{L}^2$ norm
(by
the same argument as in the case of the ordinary Bergman kernel on a
bounded
domain). We then obtain a nonzero vector
$(v_1,\dots,v_n)\in{\mathbb C}^m$ such that
$\sum_p v_p f_\alpha(z^p)=0$ for all $\alpha$. But then $\sum_p v_p
f(z^p)=0$ for all polynomials $f$ on ${\mathbb C}^m$, a contradiction.
\end{proof}
We can now show the universality of the scaling limit of the zero
correlation functions:
\begin{theo}\label{uslc} Let $(L,h)$ be a positive Hermitian line bundle
on
an
$m$-dimensional compact complex manifold $M$ with K\"ahler\ form
$\omega=\frac{i}{2}\Theta_h$, let
$\mathcal{S}=H^0(M,L^N)^k$ ($k\ge 1$), and give $\mathcal{S}$ the standard Gaussian
measure $\mu$. Then
$$\frac{1}{ N^{nk}} K_{nk}^N\left(\frac{z^1}{\sqrt{N}}, \dots,
\frac{z^n}{\sqrt{N}}\right) = K_{nkm}^\infty(z^1,\dots,z^n) +
O\left(\frac{1}{\sqrt{N}}\right)\,,$$
where $K_{nkm}^\infty(z^1,\dots,z^n)$ is given by a universal rational
function in the quantities $z^p_q, \bar z^p_q, e^{z^p\cdot \bar
z^{p'}}$,
and the error term has $\ell^{th}$ order derivatives $\le
\frac{C_{S,\ell}}{\sqrt{N}}$ on each compact subset $S\subset ({\mathbb C}^m)_n$,
for all $\ell\ge 0$. \end{theo}
\begin{proof} By taking the scaling limit of (\ref{npointcoreq}),
we obtain
\begin{equation}\label{uscleq}
K_{nkm}^\infty(z)= \frac{\mathcal{P}_{nkm}\big(\Pi^\H_1(z^p,z^{p'}),d^H_{\bar
w_{q}}\Pi^\H_1(z^p,z^{p'}), d^H_{z_{q}}\Pi^\H_1(z^p,z^{p'}),
d^H_{z_{q}}d^H_{\bar
w_{q'}}\Pi^\H_1(z^p,z^{p'})\big)}{\pi^{kn}\left[\det
\big(\Pi^\H_1(z^p,z^{p'})\big)_{1\le p,p'\le n}\right]^{k(n+1)}}\,.
\end{equation}
Indeed, since the coefficients of $\Lambda_n$ are either of degree 1 in
the coefficients of $C_n$ or of degree 2 in the coefficients of $B_n$,
we see by the proof of Theorem
\ref{npointcor}, using (\ref{dPiNH}), (\ref{dms5})--(\ref{dms6}) and
Theorem \ref{neardiag}, that the leading term of the asymptotic
expansion of
$K_{nk}^N$ is $N^{nk}$ times the right side of (\ref{uscleq}). The bound
on
the error term follows from Theorem \ref{neardiag} and Lemma
\ref{detinfty}.
Substituting into (\ref{uscleq}) the values of $\Pi^\H_1(z^p,z^{p'})$
and its
horizontal derivatives obtained from (\ref{dPiNHleft}) (with $N=1$) and
(\ref{sksl}) and canceling out common factors of $e^{-|z^p|^2/2}$ and
$\pi$,
we obtain
\begin{eqnarray}
K_{nkm}^\infty(z)&=& \frac{\mathcal{P}_{nkm}\big(e^{z^p\cdot \bar
z^{p'}},(z^{p}_q-z^{p'}_q) e^{z^p\cdot \bar z^{p'}}, (\bar z^{p'}_q
-\bar z^{p}_q)e^{z^p\cdot\bar
z^{p'}}, [(z^p_{q'}-z^{p'}_{q'})(\bar z^{p'}_q-\bar
z^{p}_q)+\delta_{qq'}]e^{z^p\cdot\bar
z^{p'}}\big)}{\pi^{kn}\left[\det
\big(e^{z^p\cdot\bar z^{p'}}\big)_{1\le p,p'\le
n}\right]^{k(n+1)}}\nonumber
\\
\label{uscleq2left}\\
&=& \frac{\mathcal{Q}_{nkm}\big(z^p_q,\bar z^p_q, e^{z^p\cdot \bar
z^{p'}}\big)}{\pi^{kn}\left[\det
\big(e^{z^p\cdot \bar z^{p'}}\big)\right]^{k(n+1)}}\,,\nonumber
\end{eqnarray} where $\mathcal{Q}_{nkm}$ is a universal polynomial
(homogeneous of
degree $k(n+1)$ in each of the variables $e^{z^p\cdot \bar z^{p'}}$ and
with
integer coefficients).
\end{proof}
\begin{rem} As we remarked previously, formula (\ref{uscleq}) is valid
for any
connection, so we can replace the left invariant vector fields with
their
right-invariant counterparts to obtain
\begin{equation}\label{uscleq2}
K_{nkm}^\infty(z)= \frac{\mathcal{P}_{nkm}\big(e^{z^p\cdot \bar
z^{p'}},z^{p}_q e^{z^p\cdot \bar z^{p'}}, \bar z^{p'}_q e^{z^p\cdot\bar
z^{p'}}, (z^p_{q'}\bar z^{p'}_q+\delta_{qq'})e^{z^p\cdot\bar
z^{p'}}\big)}{\pi^{kn}\left[\det
\big(e^{z^p\cdot\bar z^{p'}}\big)_{1\le p,p'\le n}\right]^{k(n+1)}}\,.
\end{equation}\end{rem}
\bigskip
\section {Formulas for the scaling limit zero correlation
function}\label{formulas}
We now apply the formulas from \S\S
\ref{formula-gaussian}--\ref{D-meet-S} to
transform (\ref{uscleq2}) into explicit formulas for $K_{nkm}^\infty$.
We use
the right-invariant connection $\alpha^R$ so that $d^H_{z_q} = Z_q^R.$
Indeed,
by the proofs of Theorems \ref{npointcor} and \ref{uslc} (which use
formulas
(\ref{fg6}), (\ref{fg8}), (\ref{dms4})--(\ref{dms6})), formula
(\ref{uscleq2})
becomes
\begin{equation}\label{a28}
K_{nkm}^\infty(z^1,\dots,z^n)=\frac{1}{\pi^{kn}\det A_{nkm}}\left\langle
\prod_{p=1}^n \det (a^pa^{p*})\right\rangle_{\Lambda_{nkm}}\,,
\end{equation}
where
\begin{equation}\label{a29}
\Lambda_{nkm}=C_{nkm}-B^*_{nkm}A^{-1}_{nkm}B_{nkm}
\end{equation}
with
\begin{equation}\label{a30}
\begin{array}{ll}
A_{nkm}=\left(\delta_{jj'}S(z^p,z^{p'})\right)\,,&\displaystyle
S(z,w)=\exp\left(\sum_{r=1}^m z_r\overline{w_r}\right),\\
B_{nkm}=\left(\delta_{jj'}S_{q'}(z^p,z^{p'})\right)\,,
&\displaystyle
S_{q'}(z,w)
=z_{q'}\exp\left(\sum_{r=1}^m z_r\overline{w_r}\right)\,, \\
C_{nkm}=\left(\delta_{jj'} S_{qq'}(z^p,z^{p'})\right)\,,&\displaystyle
S_{qq'}(z,w)=(\delta_{qq'}+\overline{w_q}z_{q'})
\exp\left(\sum_{r=1}^m z_r\overline{w_r}\right)\\
j,j'=1,\dots,k;\quad p,p'=1,\dots,n;&\displaystyle q,q'=1,\dots,m.
\end{array}
\end{equation}
The metric tensor $g^p$ in (\ref{fg8}) becomes a unit tensor
in the scaling limit, so there
is no $\gamma^p$ on the right in (\ref{a28}).
Because $\Pi_1^\H$ is invariant with respect to unitary transformations
and equivariant with respect to translations (i.e.,
$\Pi_1^\H(z+u,w+u)=e^{i\Im(z\cdot \bar u)} e^{-i\Im(w\cdot \bar u)}
\Pi_1^\H(z,w)$), the scaling limit zero correlation $K_{nkm}^\infty$ is
invariant with respect to the group of isometric
transformations---unitary transformations and
translations---of ${\mathbb C}^m$.
In particular, the limit one-point zero correlation, or the zero-density
function, is constant, since it is invariant under translation. Indeed
by
(\ref{a30}), $A_{1km}=e^{|z|^2}I_k$ and $\Lambda_{1km}=e^{|z|^2}I_{km}$,
where
$I_k$, resp.\ $I_{km}$, denotes the unit $k\times k$, resp.\ $(km)\times
(km)$, matrix. Thus by (\ref{a28}) and the Wick formula,
\begin{equation}\label{a32e} K_{1km}^\infty(z)=\frac{1}{\pi^k
e^{k|z|^2}}\left\langle
\det\left(\sum_{q=1}^m
\bar a_{jq}{a_{j'q}}\right)_{j,j'=1,\dots,k}
\right\rangle_{e^{|z|^2}I_{km}}
=\frac{m!}{\pi^k(m-k)!}\,.\end{equation}
Thus we define the {\it normalized n-point scaling limit zero
correlation
function} \begin{equation}\label{nslzc} \widetilde K^\infty_{nkm}(z)=
(K^\infty_{1km})^{-n}K^\infty_{nkm}(z)=\left(\frac{\pi^k(m-k)!}{m!}\right)^n
K^\infty_{nkm}(z)\,.\end{equation}
\begin{rem} These formulas also follow from \S \ref{zcorpoly}. For
example,
equation (\ref{a32e}) is a consequence of (\ref{a32b}) since
$$K_{1km}^\infty(z)=\frac{1}{N^k} K_{1k}^N(z)\,.$$
Furthermore, using the notation of \S \ref{zcorpoly}, we
observe that
\begin{equation}\label{a27}
\begin{array}{ll}
&\displaystyle\lim_{N\to\infty} S_N\left(\frac{z}{\sqrt{N}},
\frac{w}{\sqrt{N}}\right)=
\lim_{N\to\infty}\left(1+N^{-1}\sum_{r=1}^m z_r\overline{w_r}\right)^N
=S(z,w)\,,\\
&\displaystyle\lim_{N\to\infty} N^{-1/2}S_{Nq'}\left(\frac{z}{\sqrt{N}},
\frac{w}{\sqrt{N}}\right)=
S_{q'}(z,w)\,,\\
&\displaystyle\lim_{N\to\infty} N^{-1}S_{Nqq'}\left(\frac{z}{\sqrt{N}},
\frac{w}{\sqrt{N}}\right)=
S_{qq'}(z,w)\,.
\end{array}
\end{equation}
Equations (\ref{a27}) provide an alternate derivation of (\ref{a30}).
\end{rem}
\subsection{Decay of correlations}\label{decay}
Explicit formulas for the correlation functions $\widetilde K^\infty_{nkm}$ can be
obtained from (\ref{a28}), (\ref{a30}) and the Wick formula. We shall
illustrate these computations for the cases $n=2,\ k=1,2$ in \S\S
\ref{explicit1}--\ref{explicit2} below. We now note that the limit
correlations are ``short range" in the following sense:
\begin{theo}\label{shortrange} The correlation functions satisfy the estimate
$$\widetilde
K^\infty_{nkm}(z^1,\dots,z^n) = 1 +O(r^4 e^{-r^2}) \quad {\rm as}\ r\to
\infty\,, \quad r=\min_{p\ne p'}|z^p-z^{p'}|\,.$$ \end{theo}
\begin{proof} We use formula (\ref{uscleq}), which comes
from (\ref{a28})--(\ref{a29}) as in the proof of Theorem \ref{uslc}. To
determine the matrices $A,B,C$, we let $d^H_{z_q}=Z^L_q,\ d^H_{\bar
w_q}=\bar W^L_q$ (instead of the right-invariant vector fields we used
above). Recalling (\ref{dPiNHleft}), we have:
\begin{eqnarray}\label{ABC} A^{jp}_{j'p'}&=& \delta_{jj'} A^p_{p'}\,,\qquad
A^p_{p'} = \pi^m \Pi_1^\H(z^p,0;z^{p'},0)\,,\nonumber \\
B^{jp}_{j'p'q'} &=& \delta_{jj'}(z^{p}_{q'}-z^{p'}_{q'}) A^p_{p'}\,,\\
C^{jpq}_{j'p'q'} &=& \delta_{jj'}\big(\delta_{qq'}+(\bar z^{p'}_q-\bar z^{p}_{q})
( z^{p}_{q'}-z^{p'}_{q'})\big) A^p_{p'}
\,.\nonumber \end{eqnarray}
By (\ref{sksl}), \begin{eqnarray*} A^p_{p'}&=&\left\{\begin{array}{ll}1 &
p=p'\\ O(e^{-r^2/2})\, \quad & p\ne p'\end{array}\right.\,,\\
B&=&O(re^{-r^2/2})\,,\\
C&=&I+O(r^2
e^{-r^2/2})\,,\qquad C^{jpq}_{jpq}=1
\,.\end{eqnarray*}
Recalling (\ref{fg6}), we have
\begin{equation}\label{Lambda}\Lambda=I+O(r^2 e^{-r^2/2})\,,\qquad
\Lambda^{jpq}_{jpq}=1+O(r^2 e^{-r^2})\,.\end{equation} We now apply formula
(\ref{a28}); note that the Wick formula involves terms that are products
of diagonal elements of $\Lambda$, and products
that contain at least two off-diagonal elements of $\Lambda$. The former terms
are of the form $1+O(r^2e^{-r^2})$, and the latter are $O(r^4e^{-r^2})$.
Similarly, $\det A=1+O(r^4e^{-r^2})$. The desired estimate then
follows from (\ref{nslzc}).
\end{proof}
We shall see from our computations of the pair correlation below that Theorem
\ref{shortrange} is sharp. The theorem can be extended to
estimates of the connected correlation functions (called also truncated
correlation functions, cluster functions, or cumulants), as follows. The
$n$-point connected correlation function is defined as
(see, e.g., \cite[p.~286]{GJ})
\begin{equation}\label{cc1}\widetilde
T^{\infty}_{nkm}(z^1,\dots,z^n)=\sum_G(-1)^{l+1}(l-1)!
\prod_{j=1}^l \widetilde K^{\infty}_{n_jkm}(z^{i_1},\dots,z^{i_{n_j}}),
\end{equation}
where the sum is taken over all partitions $G=(G_1,\dots,G_l)$ of the
set $(1,\dots,n)$ and $G_j=(i_1,\dots,i_{n_j})$. In particular, recalling that
$\widetilde K^\infty_{1km}\equiv 1$,
\begin{eqnarray*}\label{cc2}
\widetilde T^{\infty}_{1km}(z^1)&=&\widetilde K^{\infty}_{1km}(z^1)=1\,,\\
\widetilde T^{\infty}_{2km}(z^1,z^2)&=&\widetilde K^{\infty}_{2km}(z^1,z^2)
-\widetilde K^{\infty}_{1km}(z^1)\widetilde K^{\infty}_{1km}(z^2)
\ =\ \widetilde K^{\infty}_{2km}(z^1,z^2)-1\,,\\
\widetilde T^{\infty}_{3km}(z^1,z^2,z^3)&=&\widetilde K^{\infty}_{3km}(z^1,z^2,z^3)
-\widetilde K^{\infty}_{2km}(z^1,z^2)\widetilde K^{\infty}_{1km}(z^3)
-\widetilde K^{\infty}_{2km}(z^1,z^3)\widetilde K^{\infty}_{1km}(z^2)\\
&&\ -\ \widetilde K^{\infty}_{2km}(z^2,z^3)\widetilde K^{\infty}_{1km}(z^1)
+2 \widetilde K^{\infty}_{1km}(z^1)\widetilde K^{\infty}_{1km}(z^2)
\widetilde K^{\infty}_{1km}(z^3)\\
&=&\widetilde K^{\infty}_{3km}(z^1,z^2,z^3)
-\widetilde K^{\infty}_{2km}(z^1,z^2) -\widetilde K^{\infty}_{2km}(z^1,z^3)
-\widetilde K^{\infty}_{2km}(z^2,z^3)+2\,, \end{eqnarray*}
and so on. The inverse of (\ref{cc1}) is
\begin{equation}\label{cc3}\widetilde
K^{\infty}_{nkm}(z^1,\dots,z^n)=\sum_G
\prod_{j=1}^l \widetilde T^{\infty}_{n_jkm}(z^{i_1},\dots,z^{i_{n_j}})
\end{equation}
(Moebius' theorem).
The advantage of the connected correlation functions is that they
go to zero if at least one of the distances $|z^i-z^j|$ goes to infinity (see
Corollary \ref{ccc} below).
In our case the connected correlation functions can be estimated as
follows. Define
\begin{equation}\label{cc4}
d(z^1,\dots,z^n)=\max_{\mathcal{G}}\prod_{l\in L}
|z^{i(l)}-z^{f(l)}|^2e^{-|z^{i(l)}-z^{f(l)}|^2/2}.
\end{equation}
where the maximum is taken over all oriented connected graphs
$\mathcal{G}=(V,L,i,f)$ such that $V=(z^1,\dots,z^n)$ and for every vertex $z^j\in
V$ there exist at least two edges emanating from $z^j$. Here $V$ denotes the
set of vertices of $\mathcal{G}$, $L$ the set of edges, and $i(l)$ and $f(l)$ stand
for the initial and final vertices of the edge $l$, respectively. Observe that
the maximum in (\ref{cc4}) is achieved at some graph $\mathcal{G}$, because
$te^{-t/2}\le 2/e<1$ and therefore the product in (\ref{cc4}) is less or equal
$(2/e)^{|L|}$ which goes to zero as $|L|\to\infty$.
\begin{theo}\label{cc} The connected correlation functions satisfy the estimate
$$\widetilde T^\infty_{nkm}(z^1,\dots,z^n) = O(d(z^1,\dots,z^n)) \quad {\rm as}\
\max_{p,q}|z^p-z^q|\to \infty\,,$$ provided that
$\min_{p,q}|z^p-z^q|\ge c>0$. \end{theo}
This theorem implies Theorem \ref{shortrange} because of the inversion formula
(\ref{cc3}). To prove the theorem let us remark that we can rewrite (88)
(using the Wick theorem) as a sum over Feynman diagrams. Namely,
for the normalized correlation functions $\widetilde K^\infty_{nkm}(z^1,\dots,z^n)$
we have that
\begin{equation}\label{cc5}
\widetilde K^\infty_{nkm}(z^1,\dots,z^n)
={[(m-k)!/m!]^n\over \det A_{nkm}}
\sum_{\mathcal{F}} A_{\mathcal{F}}(z^1,\dots,z^n)\,,
\end{equation}
where the sum is taken over all graphs $\mathcal{F}=(V,L,i,f)$ (Feynman diagrams)
such that $V=(z^1,\dots,z^n)$ and the edges $l\in L$ connect the paired
variables $a^{i(l)}_{jq},\; a^{f(l)*}_{j'q'}$ in a given term of the Wick sum
for $\widetilde K^\infty_{nkm}(z^1,\dots,z^n)$. The function
$A_{\mathcal{F}}(z^1,\dots,z^n)$ is a sum over all terms in the Wick sum with a
fixed Feynman diagram $\mathcal{F}$. In other words, to get
$A_{\mathcal{F}}(z^1,\dots,z^n)$ we fix pairings $(a^p_{jq},a^{p'*}_{j'q'})$
prescribed by $\mathcal{F}$ and sum up in the Wick formula over all indices $j,q$ at
every $a^p$. A remarkable property of the connected correlation functions is
that they are represented by the sum over connected Feynman diagrams (see,
e.g., \cite{GJ}),
\begin{equation}\label{cc6}\widetilde T^\infty_{nkm}(z^1,\dots,z^n)=
{[(m-k)!/m!]^n\over \det A_{nkm}}{\sum_{\mathcal{F}}}^{\rm conn}
A_{\mathcal{F}}(z^1,\dots,z^n)\,.\end{equation}
Since $\det A_{nkm}\ge c_1>0$ and $|\Lambda^{jpq}_{j'p'q'}|\le c_2<+\infty$ when
$\min_{p,q}|z^p-z^q|\ge c>0$, we conclude from (\ref{fg6}), (\ref{ABC}) and
(\ref{sksl}) that for all connected Feynman diagrams $\mathcal{F}$,
\begin{equation}\label{cc7}
A_{\mathcal{F}}(z^1,\dots,z^n)=O(d)\,,
\end{equation}
where $d=d(z^1,\dots,z^n)$ is defined in (\ref{cc4}).
Summing up over $\mathcal{F}$, we prove Theorem \ref{cc}.\qed
\begin{cor}\label{ccc} The connected correlation functions satisfy the estimate
$$\widetilde T^\infty_{nkm}(z^1,\dots,z^n) = O(R^2e^{-R^2/2}) \quad {\rm as}\
R\to\infty\,,\quad R=
\max_{p,q}|z^p-z^q|\,,$$ provided that
$\min_{p,q}|z^p-z^q|\ge c>0$. \end{cor}
\subsection {Hypersurface pair correlation}\label{explicit1} We now give
an
explicit formula [(\ref{a40})] for pair correlations in codimension 1
($k=1,\ n=2$). The case $m=1$ of this formula coincides, as it must,
with
the formula given by
\cite{H} and \cite{BBL} for the universal scaling
limit pair correlation for ${\operatorname{SU}}(2)$ polynomials. In
another paper
\cite{BSZ}, we gave a different proof of (\ref{a40}) using the
Poincar\'e-Lelong formula.
Since the scaling-limit pair correlation function
$K_{2km}^\infty(z^1,z^2)$ is invariant with respect to the group of
isometries
of
${\mathbb C}^m$, it depends only on the distance
$r=|z^1-z^2|$, so we can
set $z^1=0$ and
$z^2=(r,0,\dots,0)$. To simplify
notation, we shall henceforth write $A=A_{2km},\; B=B_{2km},\;
C=C_{2km},\;
\Lambda=\Lambda_{2km}$.
In this case, (\ref{a30}) reduces to
\begin{equation}\label{a33}
\begin{array}{lll}
A=
\left(\begin{array}{ll}
1 & 1 \\
1 & e^{r^2}
\end{array}\right);\\
B=(B^p_{p'q});&\
(B^p_{p'1})=
\left(\begin{array}{ll}
0 & 0 \\
r & re^{r^2}
\end{array}\right);&\quad
(B^p_{p'q})=
\left(\begin{array}{ll}
0 & 0 \\
0 & 0
\end{array}\right),\quad q\ge 2;\\
C=(C^{pq}_{p'q'});&\ (C^{p1}_{p'1})=
\left(\begin{array}{ll}
1 & 1 \\
1 & (1+r^2)e^{r^2}
\end{array}\right);
&\quad (C^{pq}_{p'q'})=\delta_{qq'}
\left(\begin{array}{ll}
1 & 1 \\
1 & e^{r^2}
\end{array}\right),\quad q,q'\ge 2.
\end{array}
\end{equation}
The matrix
\begin{equation}\label{a34}
\Lambda=(\Lambda^{pq}_{p'q'})=C-B^*A^{-1}B
\end{equation}
is given by
\begin{equation}\label{a35}
\Lambda^{p1}_{p'1}=
\left(\begin{array}{ll}
\displaystyle\frac{e^u-1-u}{e^u-1} & \displaystyle\frac{e^u-1-ue^u}{e^u-1} \\
\displaystyle\frac{e^u-1-ue^u}{e^u-1} & \displaystyle\frac{e^{2u}-e^u-ue^u}{e^u-1}
\end{array}\right);\quad
\Lambda^{pq}_{p'q'}=\delta_{qq'}
\left(\begin{array}{ll}
1 & 1 \\
1 & e^u
\end{array}\right),\ q,q'\ge 2\,,
\end{equation} where $u=r^2=|z^1-z^2|^2$.
By (\ref{a28}), (\ref{nslzc}) and the formula for $A$ in (\ref{a33}), we
have
\begin{equation}\label{a36}
\widetilde K^\infty_{21m}(z^1,z^2)=\frac{1}{m^2(e^u-1)}\left\langle
\left(\sum_{q=1}^m a^1_q\overline{a^1_q}\right)
\left(\sum_{q'=1}^m
a^2_{q'}\overline{a^2_{q'}}\right)\right\rangle_{\Lambda}
\end{equation}
By the Wick formula (see for example, \cite[(I.13)]{Si}),
\begin{equation}\label{a37}\begin{array}{lll}
\widetilde K^\infty_{21m}(z^1,z^2)&=&\displaystyle\frac{1}{m^2(e^u-1)}\left[
\left(\sum_{q=1}^m\langle a^1_q\overline{a^1_q}\rangle_{\Lambda_2}\right)
\left(\sum_{q'=1}^m\langle
a^2_{q'}\overline{a^2_{q'}}\rangle_{\Lambda_2}\right)
+\sum_{q,q'=1}^m \langle a^1_q\overline{a^2_{q'}}\rangle_{\Lambda_2}
\langle \overline{a^1_q}a^2_{q'}\rangle_{\Lambda_2}\right]\\[14pt]
&=&\displaystyle\frac{1}{m^2(e^u-1)}\left[
\left(\sum_{q=1}^m \Lambda^{1q}_{1q}\right)\left(\sum_{q'=1}^m
\Lambda^{2q'}_{2q'}\right)+\sum_{q,q'=1}^m \Lambda^{1q}_{2q'}
\Lambda^{2q'}_{1q}\right]\,.\end{array}\end{equation}
Substituting the values of $\Lambda^{pq}_{p'q'}$
given by (\ref{a35}), we obtain
\begin{equation}\label{a37a}
\begin{array}{lll}
\widetilde K^\infty_{21m}(z^1,z^2)&=&\displaystyle\frac{1}{m^2(e^u-1)}\left[
\left(\frac{e^u-1-u}{e^u-1}+m-1\right)
\left(\frac{e^{2u}-e^u-ue^u}{e^u-1}+(m-1)e^u\right)\right.\\[14pt]
&&\quad\displaystyle\left.+\left(\frac{e^u-1-ue^u}{e^u-1}\right)^2+(m-1)\right]\,,
\qquad u=|z^1-z^2|^2\,.\end{array}
\end{equation}
After simplification,
\begin{equation}
\widetilde K^\infty_{21m}(z^1,z^2)=\frac{u^2(e^{2u}+e^u)-2u(e^{2u}-e^u)
+m^2(e^u-1)^2e^u+m(e^u-1)^2}{m^2(e^u-1)^3}\,.
\end{equation}
Putting $u=2t$ and writing
\begin{equation}\label{a39}
\widetilde
K^\infty_{21m}(z^1,z^2)=\kappa_{1m}(|z^1-z^2|)\,,
\end{equation}
we then obtain
\begin{equation}\label{a40}
\kappa_{1m}(r)=\frac{\left[\frac{1}{2}(m^2+m)\sinh^2t+t^2\right]
\cosh t
-(m+1)t\sinh t}{m^2\sinh^3t}+\frac{(m-1)}{2m},\quad
t=\frac{r^2}{2}\,.
\end{equation}
The case $m=1$ of formula (\ref{a40}) was obtained by
Bogomolny-Bohigas-Leboeuf \cite{BBL} and Hannay \cite{H}.
As $r\to\infty$,
\begin{equation}\label{a40a} \kappa_{1m}(r)=1 +
\frac{r^4-2(m^2+1)r^2+m(3m+1)}{m^2}e^{-r^2} +O(r^4e^{-2r^2})\,.
\end{equation}
The following expansion of
the correlation function was obtained from (\ref{a40}) using Maple$^{\rm
TM}$:
\begin{eqnarray*}
\kappa_{1m}& =&\frac{m-1}{2m}t^{-1}+\frac{m-1}{2m}+
\frac{1}{6}\,{\frac {\left (m+2\right )\left (m+1\right )}{{m}^{2}}}t
-{\frac {1}{90}}\,{\frac {\left (m+4\right )\left (m+3\right )}
{{m}^{2}}}{t}^{3}\\&&\ +
{\frac {1}{945}}\,{\frac {\left (m+6\right )\left (m+5\right )}
{{m}^{2}}}{t}^{5}
-{\frac {1}{9450}}\,{\frac {\left (m+8\right )\left (m+7\right )
}{{m}^{2}}}{t}^{7}\\&&+
{\frac {1}{93555}}\,{\frac {\left (m+10\right )\left (m+9\right
)}{{m}^{2}}}{t}^{9}
-{\frac {691}{638512875}}\,{\frac {\left (m+12\right )\left (m+11
\right )}{{m}^{2}}}{t}^{11}\\&&+
{\frac {2}{18243225}}\,{\frac {\left (m+14\right )\left (m+13\right
)}{{m}^{2}}}{t}^{13}
-\cdots\;.
\end{eqnarray*}
In particular, in the one-dimensional case we have
\begin{equation}\label{a40b} \kappa_{11}(r)=\frac{1}{2}r^2 -
\frac{1}{36}r^6
+\frac{1}{720}r^{10} -\frac{1}{16800}r^{14} + \cdots\;.\end{equation}
\subsection{Pair correlation in higher codimension}\label{explicit2}
Next we compute the
two-point correlation functions for the case
$k=2$. For $k>1$, we have
\begin{equation}\label{b1} A=(A^{jp}_{j'p'})=(\delta_{jj'}A^p_{p'})\,,\quad
B=(B^{jp}_{j'p'q'})=(\delta_{jj'}B^{p}_{p'q'})\,,\quad
C=(C^{jpq}_{j'p'q'})=(\delta_{jj'}C^{pq}_{p'q'})\,,\end{equation}
where $A^p_{p'},B^{p}_{p'q'},C^{pq}_{p'q'}$ are given by (\ref{a33}).
It
follows that
\begin{equation}\label{b2}
\Lambda=(\Lambda^{jpq}_{j'p'q'})=(\delta_{jj'}\Lambda^{pq}_{p'q'})\,,\end{equation}
where $\Lambda^{pq}_{p'q'}$ is
given by (\ref{a35}).
By (\ref{a28}),
\begin{equation}\label{a41}
K^\infty_{2km}(z^1,z^2)=\frac{1}{\pi^{2k}(e^u-1)^k}\left\langle
\det\left|a_j^1\overline{a_{j'}^1}\right|_{j,j'=1,\dots,k}
\det\left|a_j^2\overline{a_{j'}^2}\right|_{j,j'=1,\dots,k}
\right\rangle_{\Lambda},\quad
a_j^p\overline{a_{j'}^{p}}=\sum_{q=1}^m
a^p_{jq}\overline{a^{p}_{j'q}}\,,
\end{equation}
where $u=r^2=|z^1-z^2|^2$ as before.
Observe that the random variables $a^p_{jq}$ and
$\overline{a^{p'}_{j'q'}}$ are
independent if either $j\not= j'$ or $q\not= q'$.
Recalling (\ref{nslzc}), we write
\begin{equation}
\widetilde K^\infty_{2km}(z^1,z^2) =\kappa_{km}(|z^1-z^2|)\,.
\end{equation}
When $k=2$, (\ref{a41}) reduces to the
following
\begin{equation}\label{a42}
\kappa_{2m}(r)=\frac
{\left\langle\left[ (a^1_1\overline{a^1_1})( a^1_2\overline{a^1_2})
-(a^1_1\overline{a^1_2})( a^1_2\overline{a^1_1})\right]
\left[ (a^2_1\overline{a^2_1})( a^2_2\overline{a^2_2})
-(a^2_1\overline{a^2_2})( a^2_2\overline{a^2_1})\right]
\right\rangle_{\Lambda}}{m^2(m-1)^2(e^u-1)^2}\,.
\end{equation}
By the Wick formula,
\begin{equation}
\kappa_{2m}(r)=\frac{d_{11}-d_{21}-d_{12}+d_{22}}{
m^2(m-1)^2(e^u-1)^2}\,,
\end{equation} where
\begin{equation}\label{a43}
\begin{array}{rl}
d_{11}= &\displaystyle\left\langle (a^1_1\overline{a^1_1})(
a^1_2\overline{a^1_2})
(a^2_1\overline{a^2_1})( a^2_2\overline{a^2_2})
\right\rangle_{\Lambda}
=\sum_{\alpha,\beta,\gamma,\delta}\left\langle a^1_{1\alpha}\overline{a^1_{1\alpha}}
a^1_{2\beta}\overline{a^1_{2\beta}} a^2_{1\gamma}\overline{ a^2_{1\gamma}}
a^2_{2\gamma}\overline{ a^2_{2\gamma}}
\right\rangle_\Lambda\\
=&\displaystyle \sum_{\alpha,\beta,\gamma,\delta}
\Lambda^{1\alpha}_{1\alpha}\Lambda^{1\beta}_{1\beta}
\Lambda^{2\gamma}_{2\gamma}\Lambda^{2\delta}_{2\delta}
\ + 2\sum_{\alpha,\beta,\gamma}
\Lambda^{1\alpha}_{1\alpha}\Lambda^{1\beta}_{2\beta}
\Lambda^{2\beta}_{1\beta}\Lambda^{2\gamma}_{2\gamma}
\ + \sum_{\alpha,\beta}
\Lambda^{1\alpha}_{2\alpha}\Lambda^{2\alpha}_{1\alpha}
\Lambda^{1\beta}_{2\beta}\Lambda^{2\beta}_{1\beta}\\
=&\left[\left(\sum_{q}\Lambda^{1q}_{1q}\right)
\left(\sum_{q}\Lambda^{2q}_{2q}\right)
\ + \sum_q \Lambda^{1q}_{2q}\Lambda^{2q}_{1q}\right]^2\,;
\end{array}
\end{equation}
similarly,
\begin{equation}\label{a43a}
\begin{array}{rl} d_{12}= &\displaystyle
\left\langle (a^1_1\overline{a^1_1})(
a^1_2\overline{a^1_2})
(a^2_1\overline{a^2_2})( a^2_2\overline{a^2_1})
\right\rangle_{\Lambda}\\
=&\left(\sum_q \left[\Lambda^{2q}_{2q}\right]^2\right)\left(\sum_q
\Lambda^{1q}_{1q}\right)^2 +
2\left(\sum_q\Lambda^{2q}_{2q}\Lambda^{2q}_{1q}\Lambda^{1q}_{2q}\right)
\left(\sum_q\Lambda^{1q}_{2q}\right) +\sum_q
\left[\Lambda^{2q}_{1q}\Lambda^{1q}_{1q}\right]^2\,,
\end{array}
\end{equation}
\begin{equation}\label{a43b}
\begin{array}{rl}
d_{21}= &\displaystyle
\left\langle (a^1_1\overline{a^1_2})( a^1_2\overline{a^1_1})
(a^2_1\overline{a^2_1})( a^2_2\overline{a^2_2})
\right\rangle_{\Lambda}\\
=&\left(\sum_q \left[\Lambda^{1q}_{1q}\right]^2\right)\left(\sum_q
\Lambda^{2q}_{2q}\right)^2 +
2\left(\sum_q\Lambda^{1q}_{1q}\Lambda^{2q}_{1q}\Lambda^{1q}_{2q}\right)
\left(\sum_q\Lambda^{2q}_{2q}\right) +\sum_q
\left[\Lambda^{2q}_{1q}\Lambda^{1q}_{2q}\right]^2\,,
\end{array}
\end{equation}
\begin{equation}\label{a43c}
\begin{array}{rl}
d_{22}= &\displaystyle
\left\langle (a^1_1\overline{a^1_2})( a^1_2\overline{a^1_1})
(a^2_1\overline{a^2_2})( a^2_2\overline{a^2_1})
\right\rangle_{\Lambda}\\
=&\left(\sum_q \left[\Lambda^{1q}_{1q}\right]^2\right)\left(\sum_q
\left[\Lambda^{2q}_{2q}\right]^2\right) +
2\sum_q\Lambda^{1q}_{1q}\Lambda^{2q}_{2q}\Lambda^{2q}_{1q}\Lambda^{1q}_{2q}
+\left(\sum_q
\Lambda^{2q}_{1q}\Lambda^{1q}_{2q}\right)^2\,.
\end{array}
\end{equation}
Substituting the values of the matrix elements of $\Lambda$ we then
obtain
\begin{equation}
\begin{array}{ll}
\kappa_{2m}(r)&= \displaystyle
\frac{(m^2-m)e^{2u}+2(m-1)e^u+2}{(e^u-1)^2m(m-1)}
-\frac{4ue^u[(m-1)e^u+1](m+1)}{(e^u-1)^3(m-1)m^2}\\
&\\
&\displaystyle +\frac {2u^2e^u[(m-1)e^{2u}+2me^u+1]}{(e^u-1)^4(m-1)m^2},\qquad
u=r^2.
\end{array}
\end{equation}
As $r\to\infty$,
\begin{equation}\label{z}
\kappa_{2m}=
1+\frac{2[r^4-2(m+1)r^2+m(m+1)]e^{-r^2}}{m^2} +O(r^4e^{-2r^2})\,.
\end{equation}
As $r\to 0$,
\begin{equation}
\begin{array}{rl}
\kappa_{2m}(r)&\displaystyle =\frac{m-2}{m}\,r^{-4}+\frac{m-2}{m}\,r^{-2}
+\frac{5m^2-7m+12}{12(m-1)m}
+\frac{(m-2)(m+2)(m+1)}{12(m-1)m^2}\,r^2\\
&\\
&\displaystyle +\frac{(m+3)(m+2)}{240(m-1)m}\,r^4
-\frac{(m-2)(m+4)(m+3)}{720(m-1)m^2}\,r^6+\dots
\end{array}
\end{equation}
When $m=2$ the asymptotics reduce to
\begin{equation}
\kappa_{22}(r)=\frac{3}{4}+\frac{r^4}{24}-\frac{r^8}{288}
+\frac{r^{12}}{4800}-\frac{r^{16}}{96768}+\dots,
\end{equation}
and in this case $\kappa_{22}$ is a series in $r^4$.
\bigskip
|
\section{Introduction}
\label{sec:intro}
The study involved in the attempt to detect gravitational waves represents
one of the most
exciting field in modern physics. The detection will provide a
decisive test of general relativity and open a new window for
astrophysical observation. Unfortunately, in spite
a great deal of effort, no one has yet succeeded in the detection
of the gravitational waves.
However ground-based laser interferometers, such as
LIGO,\cite{abramovici1992} VIRGO,\cite{bradaschia1990}
TAMA\cite{kuroda1997} and GEO600,\cite{hough1992} are being
constructed and
will soon enter a stage of data taking.
Strong sources of gravitational waves should be regions where
gravity is general relativistic and where the velocities of bulk motion
are near the speed of light. Many researchers have theoretically
investigated coalescing and colliding black holes and neutron stars as
candidates for detection. In addition, gravitational waves
emitted during stellar collapse to black holes have been
studied by perturbative calculations of a spherically symmetric
background.\cite{CPM,Seidel} \ Because of the existence of event horizon,
the emitted
gravitational waves are dominated by the quasi-normal modes of the
Schwarzschild black hole in the asymptotic region.
Some other realistic candidates exist, e.g., supernovae, rapidly
rotating neutron stars, relic gravitational waves of the early universe,
stochastic gravitational waves from cosmic strings, and so on.
In this paper we attempt to investigate whether a naked singularity, if such
exists, is a strong source of gravitational radiation,
and, moreover, to understand the dynamics and
observational meaning of naked singularity formation.
Several researchers have shown that the final fate of gravitational
collapse is not always a singularity covered by an event horizon.
For example, in the Lema\^{\i}tre-Tolman-Bondi (LTB)
spacetime,\cite{Tolman34,Bondi} a
naked shell-focusing singularity appears from generic initial data for
spherically symmetric configurations of the rest mass density and a
specific energy of the dust
fluid.\cite{Eardley,Christodoulou,Newman,Joshi-Dwivedi} \
The initial functions in the most general expandable form have been
considered.\cite{Jhingan:1997ia} \
The matter content in such a
spacetime may satisfy even the dominant energy condition.
In this case with a small disturbance of the spacetime, very short
wavelength gravitational waves, which are
created in the high density region around a singularity,
may propagate to the observer outside the dust cloud because of the
absence of an event horizon. If this is true,
extremely high energy phenomena which cannot be realized by
any high energy experiment on Earth can be observed.
In this case information regarding the physics of so-called `quantum gravity'
may be obtained. Also, these waves may be
so intense that they destroy the Cauchy horizon.
In this paper the generation of gravitational waves during the collapse of
spherical dust ball with small rotational motion is considered.
Nakamura, Shibata and Nakao\cite{nsn1993} have suggested that a
naked singularity may emit considerable gravitational wave radiation.
This was proposed from
the estimate of gravitational radiation from a
spindle-like naked singularity.
They modeled the spindle-like
naked singularity formation in gravitational collapse
by the Newtonian prolate dust collapse for a sequence of
general relativistic, momentarily static initial data. It should be noted
that the system they
considered is different from that considered in this
article and that their result is controversial. There
are numerical analyses that both support and do not support the results of
Nakamura, Shibata and
Nakao for prolate collapse\cite{Shapiro} and for cylindrical
collapse.\cite{Echeverria:1993wf,Chiba}
Due to the non-linear nature of the problem,
it is difficult to analytically solve the Einstein equation. Therefore,
numerical methods will provide the final tool. However,
its singular behavior makes
accurate numerical analysis very difficult at some stage.
In this article, we investigate odd-parity linear
gravitational waves from the collapse of an inhomogeneous
spherically symmetric dust cloud.
Even for the linearized Einstein equation, we must perform
the numerical integration.
However, in contrast to the numerical simulation of
the full Einstein equation,
high precision is guaranteed for the numerical integration of
the linearized Einstein
equation, even in the region with extremely large spacetime curvature.
Furthermore, the linear stability of known examples
of naked singularity formation is necessary
as a first step to understand the general dynamics near the ``naked
singularity formation.''
Recently, Iguchi, Nakao and Harada
\cite{Iguchi} (INH) studied odd-parity metric
perturbations around a naked singularity in the LTB spacetime. In INH,
it was found that the propagation of odd-parity gravitational waves is
not affected by collapse of a dust cloud, even if there appears a
central naked singularity. When we consider the generation of
gravitational waves from the dust collapse, we should analyze the
perturbations including their matter part. Here we investigate the
evolution equation of the odd-parity quadrupole mode for metric and matter
perturbations. This matter perturbation relates to small rotational
motion, and it produces gravitational waves during the
collapse.
We follow the derivation of the evolution equations of the perturbations
and the numerical method to integrate these equations of INH. We
investigate the time evolutions of the gauge invariant metric variables at
the symmetric center, where a naked singularity appears, and at a constant
circumferential radius $R$. We show that the gauge
invariant variable diverges only at the center and it does not propagate to
the outside.
As we know, the LTB spacetime is one candidate as a
counterexample of the cosmic censorship hypothesis (CCH), which was
introduced by Penrose.\cite{Penrose69,Penrose79} \
The CCH is a very helpful assumption, because
various theorems on the properties of black holes
were proved under this assumption.
Here the precise formulation and validity of cosmic censorship
is not our concern. We only consider the situation
in which the extremely high-density and large-curvature region
can be seen by some observer from the gravitational collapse.
Such a situation can be regarded as a naked singularity in a practical
sense since we are not yet able to predict
phenomena beyond the Planck scale.
However, here it should be noted that even from this practical
point of view, the stability of the Cauchy horizon with respect to
gravitational-wave perturbations still has an important physical
meaning. Since in the system considered here,
the mode propagating with the speed of light is the gravitational waves
only, the instability of the Cauchy horizon implies that the gravitational
wave introduces an extremely large spacetime curvature along the
null hypersurface near the Cauchy horizon from the region near the central
singularity. Hence, if the Cauchy horizon in this system is unstable,
naked singularity formation of this type might be a strong source of
gravitational radiation.
Here we comment on the problem motivated in INH, i.e.,
``nakedness of the naked singularity.''
Odd-parity matter perturbations are produced by rotational motion of
the dust cloud. Therefore the growth of the matter perturbation would
cause the centrifugal force to dominate the radial motion of the
cloud. Hence, an angular momentum bounce would occur. If this occurs at the
center, the central singularity will disappear. This situation seems to be
inevitable for the dipole mode of odd-parity matter perturbations.
In the case of a spherically symmetric system composed of
counter-rotating particles, the angular-momentum bounce can actually prevent
the naked singularity formation.\cite{Harada} \
However, for a system without spherical symmetry, it is still an open
question how the non-linear asphericity works in the final stage of
singularity formation with rotational matter perturbations because of
the difficulty of both the analytic and numerical treatment.
Further, if the initial matter perturbation is
sufficiently small, it might still be possible that the radius of the
spacetime curvature becomes the Planck length.
In this case, the behavior of other modes of perturbations
is crucial to understand the classical dynamics
in the region of the Planck scale.
As will be shown below, since there is no gravitational wave of the dipole
mode and the quadrupole mode generated in the dust cloud does not
propagate to the outside, the Cauchy horizon is marginally stable
against odd-parity perturbations originating in the aspherical
rotational motion of matter.
The paper is organized as follows: In \S \ref{sec:be}, the basic
equations are developed; in \S \ref{sec:results}, the numerical
results are presented; these results are discussed in \S
\ref{sec:discussion}; in \S \ref{sec:summary}, we summarize our
results. We adopt the geometrized units, in which $c=G=1$.
The signature of the metric tensor and
sign convention of the Riemann tensor follow
Ref.~\cite{MTW}.
\section{Basic equations}
\label{sec:be}
We consider the evolution of odd-parity perturbations of the LTB
spacetime up to linear order. The background spacetime and
perturbation method are described in INH.
\cite{Iguchi} \ In this section we briefly review them.
Using the synchronous coordinate system, the
line element of the background LTB spacetime can be expressed in the form
\begin{equation}
\label{bgmetric}
d{\bar s}^{2}= {\bar g}_{\mu\nu}dx^{\mu}dx^{\nu}\equiv
-dt^2+A^{2}(t,r)dr^{2}
+R^2(t,r)(d\theta^{2}+\sin^{2}\theta d\phi^{2}).
\end{equation}
The energy-momentum tensor for the dust fluid is
\begin{equation}
\label{bgmatter}
{\bar T}^{\mu\nu} = {\bar \rho}(t,r){\bar u}^{\mu}{\bar u}^{\nu},
\end{equation}
where ${\bar\rho}(t,r)$ is the rest mass density and
${\bar u}^{\mu}$ is the 4-velocity of the dust fluid.
In the synchronous coordinate system, the unit vector field normal to
the spacelike hypersurfaces is a geodesic, and
there is a freedom of which timelike geodesic field
is adopted as the hypersurface unit normal.
Using this freedom, we can always set ${\bar u}^{\mu}=\delta^{\mu}_{0}$,
since the 4-velocity of the spherically symmetric
dust fluid is tangent to an irrotational
timelike geodesic field.
Then the Einstein equations and the equation of motion for the
dust fluid reduce to the simple equations
\begin{eqnarray}
A &=& \frac{\partial_{r}R}{\sqrt{1+f(r)}}, \label{eq:A} \\
{\bar \rho}(t,r) &=& \frac{1}{8\pi}
\frac{1}{R^2 \partial_{r}R}{dF(r)\over dr},\label{eq:einstein} \\
(\partial_{t}R)^2-\frac{F(r)}{R} &=& f(r),\label{eq:energyeq}
\end{eqnarray}
where $f(r)$ and $F(r)$ are arbitrary functions of the radial
coordinate, $r$. From Eq.\ (\ref{eq:einstein}), $F(r)$ is related
to the Misner-Sharp mass function,\cite{Misner} $m(r)$,
of the dust cloud in the manner
\begin{equation}
\label{mass}
m(r) = 4\pi \int_0^{R(t,r)}{\bar \rho}(t,r)R^2dR = 4\pi
\int_0^r{\bar \rho}(t,r)R^2\partial_{r}R dr =\frac{F(r)}{2}.
\end{equation}
Hence Eq.\ (\ref{eq:energyeq}) might be
regarded as the energy equation per unit mass.
This means that
the other arbitrary function, $f(r)$, is recognized as
the specific energy of the dust fluid.
The motion of the dust cloud is completely specified
by the function $F(r)$
(or equivalently, the initial distribution of
the rest mass density, ${\bar \rho}$) and the specific energy, $f(r)$.
When we restrict our calculation to the case that the symmetric center,
$r=0$, is initially regular, the central shell focusing singularity
is naked if and only if
$\partial_{r}^{2}{\bar\rho}|_{r=0}<0$ is initially
satisfied for the marginally bound collapse, $f(r)=0$.
\cite{Singh-Joshi,Jhingan} \
For a collapse that is not marginally bound, there exists a similar
condition as an inequality for a value depending on the functional forms
of $F(r)$ and $f(r)$.\cite{Newman,Singh-Joshi,Jhingan} \
Let us now derive the perturbation equations.
The perturbed metric tensor is expressed in the form
\begin{equation}
\label{metric}
g_{\mu\nu} = {\bar g}_{\mu\nu}+h_{\mu\nu},
\end{equation}
where ${\bar g}_{\mu\nu}$ is a metric tensor of a
spherically symmetric background spacetime
and $h_{\mu\nu}$ is a perturbation.
The energy-momentum tensor is written in the form
\begin{equation}
\label{ene-mom}
T_{\mu\nu} = \bar{T}_{\mu\nu} + \delta T_{\mu\nu},
\end{equation}
where $\bar{T}_{\mu\nu}$ is a background quantity and
$\delta T_{\mu\nu}$ is a perturbation.
By virtue of the spherical symmetry of the background spacetime,
${\bar T}_{\mu\nu}$ can be expressed in the form
\begin{eqnarray}
\label{GS-matter}
\bar{T}_{\mu\nu}dx^{\mu}dx^{\nu} &=&
\bar{T}_{ab}dx^adx^b+\frac{1}{2}\bar{T}_B^{~B}R^2(t,r)d\Omega^2,
\end{eqnarray}
where the subscripts and superscripts $a,b,\cdots$ represent $t$ and $r$,
while $A,B,\cdots$ represent $\theta$ and $\phi$.
The odd-parity perturbations of
$h_{\mu\nu}$ and $\delta T_{\mu\nu}$ are expressed
in the form
\begin{eqnarray}
\label{GS-pmetric}
h_{\mu\nu} &=& \left(\begin{array}{ccc}
0 & 0 & h_0(t,r)\Phi^m_{l~B} \\
& 0 & h_1(t,r)\Phi^m_{l~B} \\
\mbox{sym} & & h_2(t,r)\chi^m_{l~AB}
\end{array} \right),
\end{eqnarray}
\begin{eqnarray}
\label{GS-pmatter}
\delta T_{\mu\nu} &=& \left(\begin{array}{ccc}
0 & 0 & t_0(t,r)\Phi^m_{l~B} \\
& 0 & t_1(t,r)\Phi^m_{l~B} \\
\mbox{sym} & & t_2(t,r)\chi^m_{l~AB}
\end{array} \right) ,
\end{eqnarray}
where $\Phi^m_{l~B}$ and $\chi^m_{l~AB}$ are odd-parity vector and
tensor harmonics associated with the spherical symmetry of
the background spacetime.\cite{Regge} \ We set all the arbitrary
constants in the definitions of harmonics to unity.
We then introduce the gauge-invariant
variables defined by Gerlach and Sengupta.\cite{Gerlach-Sengupta} \
The metric variables are given by
\begin{equation}
\label{ka}
k_a = h_a-\frac{1}{2}R^2\partial_a\left(\frac{h_2}{R^2}\right).
\end{equation}
The matter variables are given by the combinations
\begin{eqnarray}
\label{la}
L_a &=& t_a-\frac{1}{2}\bar{T}_B^{~B}h_a ,\\
\label{l}
L &=& t_2-\frac{1}{2}\bar{T}_B^{~B}h_2 .
\end{eqnarray}
In the LTB case, the odd-parity gauge-invariant
matter variables become
\begin{eqnarray}
\label{La-L}
L_0 = \bar{\rho}(t,r) U(t,r)~~~~~~{\rm and}~~~~~~ L_1 = L = 0,
\end{eqnarray}
where $U(t,r)$ represents the perturbation of the 4-velocity as
$\delta u_\mu = (0,0,U(t,r)\Phi_{l~B}^m)$.
The evolution equation for the matter variable (Eq.\ (3$\cdot$19) in INH),
\begin{equation}
\label{T-Bconserve}
\partial_t\left(AR^2L_0\right) = 0,
\end{equation}
is easily integrated, and we obtain
\begin{equation}
\label{kakuundouryou}
L_0 = \frac{1}{AR^2}{dJ(r)\over dr},
\end{equation}
where $J(r)$ is an arbitrary function depending only on
$r$. From Eqs.\ (\ref{eq:einstein}), (\ref{La-L}), and (\ref{kakuundouryou}),
we obtain the relation
\begin{eqnarray}
\label{U(r)}
U(t,r)&=&8\pi \sqrt{1+f(r)}\frac{dJ(r)/dr}{dF(r)/dr} \\
&\equiv& U(r),
\end{eqnarray}
so $U(t,r)$ is independent of the time coordinate $t$.
We introduce a gauge-invariant variable for the metric as
\begin{equation}
\label{psis}
\psi_s \equiv {1\over A}\left[\partial_t
\left(\frac{k_1}{R^2}\right)-\partial_r
\left(\frac{k_0}{R^2}\right)\right].
\end{equation}
The metric perturbation variables, $k_0$ and $k_1$, are reconstructed
from the linearized Einstein equations,
\begin{eqnarray}
\label{wave2}
\partial_r\left(R^4\psi_s\right)+A\left(l-1\right)
\left(l+2\right)k_0 &=& 16\pi AR^2L_0, \\
\label{wave3}
\partial_t\left(R^4\psi_s\right)+\frac{1}{A}
\left(l-1\right)\left(l+2\right)k_1 &=& 0,
\end{eqnarray}
by substituting $\psi_s$. It was
shown in INH that $\psi_s$ is closely connected to the tetrad components
of the magnetic part of the Weyl tensor. These components diverge where
and only where $\psi_s$ diverges. From the linearized Einstein equations
we obtain the linearized evolution equation for the odd-parity
perturbation as
\begin{eqnarray}
\label{wave4}
\partial_{t}\left(\frac{A}{R^2}\partial_{t}
\left(R^4\psi_s
\right)\right)&-&\partial_{r}
\left(\frac{1}{AR^2}\partial_{r}
\left(R^4\psi_s\right)\right) +
\left(l-1\right)
\left(l+2\right)A\psi_s \nonumber\\
&=& -16\pi\partial_{r}\left(\frac{1}{AR^2}{dJ\over dr}\right).
\end{eqnarray}
The regularity conditions at the center are also considered in INH.
The result for the matter perturbation $L_0$ is given by
\begin{equation}
\label{r-0-L}
L_0 \longrightarrow L_c(t)r^{l+1}+O(r^{l+3}).
\end{equation}
Therefore the matter perturbation variables vanish at the regular center
independent of the value of $l$.
The metric variable $\psi_s$ behaves near the center as
\begin{eqnarray}
\label{r-0-psis}
\psi_{s} &\longrightarrow&
\psi_{sc}(t)r^{l-2}+O(r^l)
~~~~~~~~~~~~\mbox{for $l\geq 2$}, \\
\label{r-0-psis1}
\psi_{s} &\longrightarrow& \psi_{sc}(t)r+O(r^3)
~~~~~~~~~~~~~~~~\mbox{for $l=1$}.
\end{eqnarray}
From the above equations, it is shown that only the quadrupole mode,
$l=2$, of $\psi_{s}$ does not vanish at the regular center.
\section{Results}
\label{sec:results}
We numerically solve the wave equation (\ref{wave4}) in the case of
marginally bound collapse, $f(r)=0$, and the quadrupole mode, $l=2$.
We follow numerical techniques employed in INH to integrate the time
evolution of the
perturbations. The numerical code is essentially the same as that in INH,
which is tested by comparison with the analytic solution for the Minkowski
background.
By virtue of the relation $f(r)=0$,
we can easily integrate Eq.\ (\ref{eq:energyeq}) and obtain
\begin{equation}
\label{f=0}
R(t,r) = \left(\frac{9F}{4}\right)^{1/3}[t_0(r)-t]^{2/3},
\end{equation}
where $t_0(r)$ is an arbitrary function of
$r$. The naked singularity formation time is $t_0=t_0(0)$. Using the freedom
for the scaling of $r$, we choose $R(0,r)=r$. This scaling of $r$
corresponds to the following choice of $t_{0}(r)$:
\begin{equation}
\label{t0}
t_0(r) = \frac{2}{3\sqrt{F}}r^{3/2}.
\end{equation}
Here note that, from Eq.\ (\ref{eq:A}), the background
metric variable, $A$, is equal to $\partial_{r}R$.
Then, the wave equation (\ref{wave4}) becomes
\begin{eqnarray}
\label{wave-eq}
\frac{\partial^2 \psi_s}{\partial t^2}-\frac{1}{(\partial_{r}R)^2}
\frac{\partial^2 \psi_ s}{\partial r^2}
&=& \frac{1}{(\partial_{r}R)^{2}}\left(6\frac{\partial_{r}R}{R}
-\frac{\partial_{r}^{2}R}{\partial_{r}R}\right)\frac{\partial
\psi_s}{\partial r}-\left(6\frac{\partial_{t}R}{R}
+\frac{\partial_{t}\partial_{r}R}{\partial_{r}R}\right)
\frac{\partial
\psi_s}{\partial t} \nonumber \\
& &-4\left[\left(\frac{\partial_{t}\partial_{r}R}{\partial_{r}R}\right)
\frac{\partial_{t}R}{R}
+\frac{1}{2}\left(\frac{\partial_{t}R}{R}\right)^2\right]\psi_s \nonumber \\
& &-\frac{16\pi}{(\partial_r R)R^2}\partial_{r}
\left(\frac{r^2 \rho (r) U(r)}{(\partial_r R)R^2}\right),
\end{eqnarray}
where $\rho (r) = \bar{\rho}(0,r)$ is the density profile at $t=0$.
We solve this partial differential equation numerically.
Before getting into the detailed explanation of the numerical techniques, we
comment on the behavior of the matter perturbation variable $L_0$ around
a naked singularity on the slice $t=t_0$.
The regularity conditions of $L_0$ and $\bar{\rho}$
determine the behavior of $U(r)$ near the center as
\begin{equation}
\label{dJdr-c}
U(r) \propto r^{l+1}.
\end{equation}
This property does not change even if a central singularity
appears. However, the $r$ dependence of $R$ and $A$ near the center changes at
that time. Assuming a rest mass density profile of the form
\begin{equation}
\rho(r) = \rho_0 + \rho_n r^n + \cdots ,
\end{equation}
we obtain the relation
\begin{equation}
t_0(r) \propto t_0 + t_n r^n
\end{equation}
from Eqs.\ (\ref{eq:einstein}) and (\ref{t0}),
where $n$ is a positive even integer. After substituting this
relation into Eq.\ (\ref{f=0}),
the lowest order term is absent from the square brackets of it.
Then we obtain the behavior of $R$ and $A$ around the central singularity as
\begin{equation}
\label{singr}
R(t_0,r) \propto r^{1+ {2\over3} n},
\end{equation}
and
\begin{equation}
\label{singa}
A(t_0,r) \propto r^{{2\over3} n}
\end{equation}
on the slice $t=t_0$.
As a result, we obtain the $r$
dependence of $L_0$ around the center when the naked singularity appears as
\begin{equation}
L_0(t_0,r) \propto r^{l-2n+1}.
\end{equation}
For example, if $l=2$ and $n=2$, then $L_0$ is inversely proportional to $r$
and diverges toward the central naked singularity. Therefore the source term
of the wave equation is expected to have a large magnitude around the naked
singularity. Thus the metric perturbation variable $\psi_s$ as well as
matter variable $L_0$ may diverge toward the naked singularity.
Instead of the $(t,r)$ coordinate system, we
introduce a single-null coordinate system, $(u,r')$,
where $u$ is an outgoing null coordinate and chosen so that it
agrees with $t$ at the symmetric center and we choose $r'=r$.
We perform the numerical integration along two characteristic
directions. Therefore we use a double null grid in the numerical
calculation. A detailed explanation of these coordinates is given in
INH. By using this new coordinate system, $(u,r')$, Eq.\ (\ref{wave-eq})
is expressed in the form
\begin{eqnarray}
\label{dphis/dlambda}
\frac{d\phi_s}{du} &=&
-\frac{\alpha}{R}\left[3\partial_{r}R
+{1\over2}R(\partial_{t}R)\partial_{t}\partial_{r}R
-\frac{5}{4}(\partial_{t}R)^{2}\partial_{r}R\right]\psi_s \nonumber \\
& &
-\frac{\alpha}{2}\left[\frac{\partial_{r}^{2}R}{(\partial_{r}R)^2}
-\frac{2}{R}\left(1-\partial_{t}R\right)
\right]\phi_s -\frac{8\pi \alpha}{R}\partial_{r}\left(
\frac{r^2 \rho (r) U(r)}{(\partial_rR)R^2}\right),\\
\label{delpsi/delrprime}
\partial_{r'}\psi_{s}&=& \frac{1}{R}\phi_s
-3\frac{\partial_{r}R}{R}
\left(1+\partial_{t}R\right)\psi_s,
\end{eqnarray}
where the ordinary derivative on the left-hand side of
Eq.\ (\ref{dphis/dlambda}) and the partial derivative on the left-hand
side of Eq.\ (\ref{delpsi/delrprime}) are given by
\begin{eqnarray}
\frac{d}{du} &=& \partial_u +
\frac{dr'}{du}\partial_{r'}
= \partial_u-\frac{\alpha}{2\partial_r R}\partial_{r'} \nonumber \\
&=&\frac{\alpha}{2}\partial_t-\frac{\alpha}{2\partial_r R}\partial_{r},\\
\label{r'a}
\partial_{r'} &=& -\frac{(\partial_r u)_t}{(\partial_t
u)_r}\partial_{t}+\partial_{r}=
(\partial_{r}R)\partial_{t}+\partial_{r},
\end{eqnarray}
respectively. Also, $\phi_s$ is defined by
Eq.\ (\ref{delpsi/delrprime}) and $\alpha$ is given by
\begin{equation}
\alpha\equiv {1\over (\partial_{t} u)_r}.
\end{equation}
We integrate Eq.\ (\ref{dphis/dlambda}) using the scheme of an explicit first
order difference equation. We use the trapezoidal rule,
\begin{equation}
\psi_{s j+1}=\psi_{s j}+\frac{\Delta r'}{2}\left((\partial_{r'}\psi_{s})_{j}+(\partial_{r'}\psi_{s})_{j+1}\right),
\end{equation}
to integrate Eq.\ (\ref{delpsi/delrprime}).
For the boundary condition at the center we demand that $\psi_s$ behaves
as $\psi_{sc}(t)+\psi_{s2}(t)r^2$ on a surface of $t=\mbox{const}$. We
numerically realize this condition by two-step interpolation. First
the values of $\psi_s$ are derived at two points on the surface of
$t=\mbox{const}$ from the interpolation on the slices of $u=\mbox{const}$.
Next, using
these two values, the central value of $\psi_s$ is derived from the
interpolation on the slice of $t=\mbox{const}$.
Another way to determine the central value of
$\psi_s$ is as follows. We first obtain the central value of $\phi_s$
from Eq.\ (\ref{dphis/dlambda}). From Eq.\ (\ref{delpsi/delrprime}) and
the boundary conditions, the relation of $\psi_s$ and $\phi_s$ at the
center is given by
\begin{equation}
\label{phi-psi}
\phi_s = 3 \partial_{r'}R \psi_s.
\end{equation}
Using this relation the central value of $\psi_s$ is obtained. In our
numerical analyses the results of these two methods agree well.
We assume $\psi_s$ vanishes on the initial null hypersurface. Therefore,
there
exist initial ingoing waves which offset the waves produced by the
source term on the initial null hypersurface. In INH, it is confirmed
that this type of the initial ingoing waves propagate through the dust
cloud without net
amplification even when they pass through the cloud just before the
appearance of the naked singularity. Therefore those parts of the metric
perturbations would not diverge at the center.
We adopt the initial rest mass density profile
\begin{equation}
\label{density}
\rho(r)=\rho_0 \frac{1+ \exp\left(-\frac{1}{2}\frac{r_1}{r_2}\right)}
{1+ \exp\left(\frac{r^n -r_1^n}{2 r_1^{n-1}r_2}\right)},
\end{equation}
where $\rho_0$, $r_1$ and $r_2$ are positive constants and $n$ is a
positive even
integer. As a result the dust fluid spreads all over the space. However,
if $r \gg r_1,r_2$, then $\rho(r)$ decreases exponentially, so that the dust
cloud is divided into the core part and the envelope which would be
considered as the vacuum region essentially. We define a core radius as
\begin{equation}
\label{cradious}
r_{\mbox{\scriptsize core}}=r_1+\frac{r_2}{2}.
\end{equation}
If we set $n=2$, there appears a central naked
singularity. This singularity becomes locally or globally naked depending
on the parameters ($\rho_0,r_1,r_2$). However, if the integer
$n$ is greater than $2$, the final state of the dust cloud is a black
hole independently of the parameters. Then we consider three different
density profiles connected with three types of the final state of the
dust cloud, globally and locally naked singularities and a black
hole. The outgoing null coordinate $u$ is chosen so that it agrees
with the proper time at the symmetric center. Therefore, even if the
black hole background is considered, we can analyze the inside of the
event horizon.
Corresponding parameters are given in Table \ref{tab:parameter}.
Using this density profile, we numerically calculate the total
gravitational mass of
the dust cloud $M$. In our calculation we adopt the total
mass $M$ as the unit of the variables.
The source term of Eq.\ (\ref{dphis/dlambda}),
\begin{equation}
\label{souce}
S(t,r)=-\frac{8\pi \alpha}{R}\partial_{r}\left(
\frac{r^2 \rho (r) U(r)}{(\partial_rR)R^2}\right),
\end{equation}
is determined by $U(r)$. As mentioned above, the constraints on the
functional form of
$U(r)$ are given by the regularity condition of $L_0$. From
Eq.\ (\ref{dJdr-c}), $U(r)$ should be
proportional to $r^{l+1}$ toward the center.
We localize the matter perturbation near the center to diminish the
effects of the initial ingoing waves. Therefore we define $U(r)$ such that
\begin{equation}
\label{dJdr}
r^2 \rho(r) U(r) = \left\{
\begin{array}{ccc}
U_0 \left( \frac{r}{r_b} \right)^5 \left(1-\left
( \frac{r}{r_b} \right)^2 \right)^5 &{\mbox{for}} & 0\leq r\leq r_b,\\
0 &{\mbox{for}} & r>r_b,
\end{array}
\right.
\end{equation}
where $U_0$ and $r_b$ are arbitrary constants. In our numerical calculation
we chose $r_b$ as $r_{\mbox{\scriptsize core}}/2$. This choice of $r_b$ has no
special meaning, and the results of our numerical calculations are not
sensitive to it.
First we observe the behavior of $\psi_s$ at the center. The results are
plotted in Fig.\ \ref{fig:center}. The initial oscillations correspond to
the initial ingoing waves. After these oscillations, $\psi_s$ grows
proportional to $(t_0-t)^{- \delta}$ for the naked singularity cases
near the formation epoch of the naked singularity.
For the case of black hole formation, $\psi_s$ exhibits power-law
growth in the early part. Later its slope gradually changes
but it grows faster than in the case of the naked singularity.
For the naked singularity cases, the power-law indices $\delta$ are
determined by $(t_0-t)\dot{\psi_s}/\psi_s$ locally. The results are
shown in Fig.\ \ref{fig:index}. From this figure we read the final
indices as 5/3 for both naked cases.
Therefore the metric perturbations diverge at the central naked singularity.
\begin{table}
\caption{Parameters of initial density profiles, power law indices, and
damped oscillation frequencies.}
\label{tab:parameter}
\begin{tabular}{cccccccc}
& final state & $\rho_0$ & $r_1$ & $r_2$ & $n$ &power index&
damped oscillation frequency\\ \hline
(a) & globally naked & $1 \times 10^{-2}$ & 0.25 & 0.5 & 2 & 5/3 & --- \\
(b) & locally naked & $1 \times 10^{-1}$ & 0.25 & 0.5 & 2 & 5/3 & 0.37+0.089$i$\\
(c) & black hole & $2 \times 10^{-2}$ & 2 & 0.4 & 4 & --- & 0.37+0.089$i$\\
\end{tabular}
\end{table}
We also observe the wave form of $\psi_s$ along the line of a
constant circumferential radius outside the dust cloud. The results are
shown in Figs.\ \ref{fig:outsideg}--\ref{fig:outsideb}. Figure
\ref{fig:outsideg} displays the wave form of the globally naked case
(a), Fig.\ \ref{fig:outsidel} displays the wave form of the
locally naked case (b), and Fig.\ \ref{fig:outsideb} displays the
wave form of the black hole case (c).
The initial oscillations
correspond to the initial ingoing waves. In the case of a locally naked
singularity and black hole formation, damped oscillations dominate the
gravitational waves. We read the frequencies and
damping rates of these damped oscillations from Figs.\
\ref{fig:outsidel} and \ref{fig:outsideb} and give them in terms of
complex frequencies as $0.37+0.089i$ for locally naked
and black hole cases. These agree well with the
fundamental quasi-normal frequency of the quadrupole mode
$(2M\omega = 0.74734 + 0.17792 i)$ of a Schwarzschild black hole given
by Chandrasekhar and Detweiler.\cite{Chandra} \ In the globally naked
singularity case (a), we did not see this damped oscillation
because of the
existence of the Cauchy horizon. In all cases the gravitational waves
generated by matter perturbations are at most quasi-normal modes of a
black hole, which is generated outside the dust cloud.
Therefore intense odd-parity gravitational waves would not be
produced by the inhomogeneous dust cloud collapse. We should not expect
that the central extremely high density region can be observed by this
mode of gravitational waves.
We can calculate the radiated power of the gravitational waves and thereby
grasp the physical meaning of the gauge-invariant
quantities.\cite{CPM} \ To relate the perturbation of the metric to the
radiated
gravitational power, it is useful to specialize to the radiation gauge,
in which the tetrad components $h_{(\theta)(\theta)}-h_{(\phi)(\phi)}$
and $h_{(\theta)(\phi)}$ fall off as $O(1/R)$, and all other tetrad
components fall off as $O(1/R^2)$ or faster. Note that, in vacuum at
large distance, the spherically symmetric background metric is given by
the Schwarzschild solution, where hereafter we adopt the Schwarzschild
coordinates,
\begin{equation}
\label{Schmetric}
ds^2 = -\left( 1-\frac{2M}{R}\right)d\tau ^2 +\left
( 1-\frac{2M}{R}\right)^{-1} dR^2 + R^2 \left(
d\theta^{2}+\sin^{2}\theta d\phi^{2}\right) .
\end{equation}
The relation between the line elements Eq.\ (\ref{bgmetric}) and
Eq.\ (\ref{Schmetric}) is given by the transfer matrix
\begin{eqnarray}
d\tau &=& \frac{1}{1-(\partial_t R)^2} (dt+\partial_r R \partial_t R
dr), \\
dR &=& \partial_t Rdt + \partial_r R dr.
\end{eqnarray}
In this gauge, the metric perturbations in Eq. (\ref{GS-pmetric}) behave
as
\begin{eqnarray}
h_0, h_1 &=& O\left(\frac{1}{R}\right), \\
h_2 &=& w(\tau - R_*) + O(1),
\end{eqnarray}
where
\begin{equation}
R_* = R + 2M \ln \left( \frac{R}{2M} - 1\right) + \mbox{const}.
\end{equation}
Then, the gauge-invariant metric perturbations (\ref{ka}) are calculated
as
\begin{eqnarray}
k_0 &=& -\frac{1}{2}w^{(1)}R + O(1), \\
k_1 &=& \frac{1}{2}w^{(1)}R + O(1),
\end{eqnarray}
where $w^{(1)}$ denotes the first derivative of $w$
with respect to its argument.
In this radiation gauge, the radiated power $P$ per unit solid angle is
given by the formula which was derived by Landau and Lifshitz
\cite{Landau} from their stress-energy pseudo-tensor:
\begin{equation}
\frac{dP}{d\Omega}=\frac{R^2}{16\pi}\left[\left(\frac{\partial
h_{(\theta)(\phi)}}{\partial \tau}\right)^2
+\frac{1}{4}\left(\frac{\partial h_{(\theta)(\theta)}}{\partial
\tau}-\frac{\partial h_{(\phi)(\phi)}}{\partial \tau}\right)^2\right].
\end{equation}
For the axisymmetric mode, i.e., $m=0$, the above formula is reduced to
\begin{equation}
\frac{dP}{d\Omega}=\frac{1}{64\pi}(w^{(1)})^2A_l (\theta),
\end{equation}
where
\begin{equation}
A_l (\theta) \equiv \frac{2l+1}{4\pi}\sin^4\theta\left(\frac{d^2P_l(\cos\theta)}{(d\cos\theta)^2}\right)^2.
\end{equation}
Then, by using the gauge-invariant quantities and integrating over the
all solid angles, the formula for the power of gravitational radiation
is obtained in the following form:
\begin{eqnarray}
\frac{dP}{d\Omega}&=&\frac{1}{16\pi}\frac{k_0^2}{R^2}A_l
(\theta)=\frac{1}{16\pi}\frac{k_1^2}{R^2}A_l (\theta),\\
P &=& \frac{1}{16\pi} B_l \frac{k_0^2}{R^2} = \frac{1}{16\pi} B_l
\frac{k_1^2}{R^2} ,
\end{eqnarray}
where
\begin{equation}
B_l \equiv \frac{(l+2)!}{(l-2)!}.
\end{equation}
Using Eq. (\ref{wave3}), the radiated power $P$ of the quadrupole mode
is given by
\begin{equation}
P = \frac{3}{32\pi}R'^2\left[\partial_{\tau}\left(R^3 \psi_s\right)\right]^2.
\end{equation}
Figure \ref{fig:power} displays the time evolution of the radiated power
$P$. The radiated power also has a finite value at the
Cauchy horizon. The total energy radiated by odd-parity quadrupole
gravitational waves during the dust collapse should not diverge.
\section{Discussion}
\label{sec:discussion}
First we consider the behavior of the source term $S(t,r)$ around the
naked singularity. From the regularity conditions and
Eqs.\ (\ref{dJdr-c}), (\ref{singr}), and (\ref{singa}),
the asymptotic behavior of the source term is obtained as
\begin{equation}
S(t,r) \propto r^{l-1}
\end{equation}
for $t<t_0$ and
\begin{equation}
S(t,r) \propto r^{l-\frac{8}{3}n-1},
\end{equation}
at $t=t_0$. For example,
in the case $l=2$ and $n=2$, the souce term
behaves on $t=t_0$ as
\begin{equation}
S(t,r) \propto r^{-13/3},
\end{equation}
and then it diverges at the center. Thus the divergency of $\psi_s$ at the
center originates from the source term. To confirm this,
we numerically integrate the source term along the ingoing null lines
with respect to $u$ and estimate the central value of $\phi_s$. We define this
`estimated' value as
\begin{equation}
\Phi_s \equiv \int S(t,r) du.
\end{equation}
Using Eq.\ (\ref{phi-psi}) we can define the estimated value of $\psi_s$ as
\begin{equation}
\Psi_s \equiv \frac{\Phi_s}{3 \partial_rR}.
\end{equation}
We plot it in Fig.\ \ref{fig:int} together with the corresponding
$\psi_s$. The estimated value has the same power-law index of $\psi_s$.
We conclude that the behavior of $\psi_s$ is determined
by the source term in the dust cloud.
We next consider the stability of the Cauchy horizon.
We found that the metric perturbation produced by the source term
does not propagate outside the dust cloud, except for quasi-normal ringing.
The source term, which controls $\psi_s$, does not diverge at the Cauchy
horizon. Therefore $\psi_s$ should not diverge at the Cauchy horizon and
should not
destroy it. Then, even if odd-parity perturbations are considered, it
will not be the case that the LTB spacetime
loses its character as
a counterexample to
CCH due to Cauchy horizon instability. Also, it does not seem that such
collapse is a strong source of gravitational waves.
In this paper, we have dealt with the marginally bound
case. For the case of non-marginally bound collapse,
the condition of the appearance
of the central naked singularity is slightly different from that in
the above
case\cite{Singh-Joshi,Jhingan} and hence there is the possibility
that the behavior of $\psi_{s}$ in this case is different from that in the
marginally bound case. However, it is well known that
the limiting behavior of the metric as
$t \rightarrow t_0(r)$ is common for all the cases:\cite{Landau}
\begin{equation}
R \approx \left(\frac{9F}{4}\right)^{1/3}\left(t_0-t\right)^{2/3}
,~~~A \approx \left(\frac{2F}{3}\right)^{1/3} \frac{t_0'}{\sqrt{1+f}}
\left(t_0-t\right)^{-1/3}.
\end{equation}
We conjecture that the results of the
perturbed analysis for the non-marginal collapse would be similar to
the results for the marginal bound.
\section{Summary}
\label{sec:summary}
We have studied the behavior of the odd-parity perturbation
in the LTB spacetime including the matter perturbation.
For the quadrupole mode, where gravitational waves exist, we
have numerically investigated the wave equation.
For the case of naked singularity formation,
the gauge-invariant metric variable, $\psi_s$, diverges according to a power
law with power index 5/3 at the center. This power index is closely
related to the behavior of the matter perturbation around the
center. We have also observed $\psi_s$ at a
constant circumferential radius. For the globally naked case, we cannot
see intense gravitational waves propagated from the center just before
the crossing of the Cauchy horizon. For the locally naked case, we
have confirmed that there exist quasi-normal oscillations. As a
result, we conclude that the type of singularity changes
due to the odd-parity perturbation because $\psi_s$ diverges at the
center. However, the Cauchy horizon is marginally stable against
odd-parity perturbations.
At the final stage of the collapse, the effects of the rotational motion
are important and the centrifugal force might dominate the radial motion.
If this is true, the central singularity would disappear when an
odd-parity matter perturbation is introduced.
For the dipole mode, such a situation seems to be inevitable.
However, we should note that it is a non-trivial and open question
how non-linear asphericity affects the final fate of the
singularity-formation process.
Further, in the case of initially sufficiently small aspherical
perturbations, the radius of spacetime curvature at the center might
reach the Planck length, and hence there is still the possibility that
the naked singularity is formed there in a practical sense.
However, as our present analysis has revealed, since the Cauchy horizon
is stable with respect to odd-parity linear perturbations,
there is little possibility that this collapse is a strong source
of odd-parity gravitational waves.
There remains important related works to be completed.
The first problem is to consider the even-parity mode in which the
metric and matter perturbations are essentially coupled with each other.
We are now investigating this problem.
Finally, we should consider the non-linear effects to complete this analysis.
This problem will be analyzed in the future.
\section*{Acknowledgements}
We would like to thank T. Nakamura for helpful and useful discussions and
N. Sugiyama for careful reading of the manuscript. We are also
grateful to H. Sato and colleagues in the theoretical astrophysics group
at Kyoto University for useful comments and encouragement.
This work was partly supported by Grants-in-Aid for Scientific
Research (No.\ 9204) and for Creative Basic Research (No.\ 09NP0801)
from the Japanese Ministry of Education,
Science, Sports and Culture.
|
\section{Introduction}
It has been observed that many galaxies and clusters of galaxies are endowed
with a magnetic field with typical strength of order $10^{-6}$ \rm G \cite{kron}.
The origin of
these large scale magnetic fields is unknown. In order for magnetic fields to
have this order of magnitude it is widely believed that an enormous
amplification of an initial seed field must have taken place. This amplification is usually
explained by dynamo theory which can enhance the magnetic field exponentially \cite{dyn}.
However, dynamos cannot create a magnetic field, and so in order for them to act
they require a seed. At present it is not clear whether this seed
field has its origin from some astrophysical mechanism after recombination,
during the epoch of galaxy formation and afterwards or whether the seed field is
of primordial origin, produced in the very early universe. In the latter
scenario it is believed that the primordial magnetic field would have been frozen
into the highly conductive plasma as the universe expanded and cooled. Because
of the high conductivity diffusion would be small and magnetic flux conserved.
If a magnetic field was produced in the early universe and was present at the
time of recombination it may have had a significant effect on many astrophysical
processes including the formation of galaxies and stars.
There are several ways of obtaining limits on cosmological magnetic fields.
Limits have been obtained by Faraday rotation measurements of intergalactic
fields \cite{vallee,bla-bur-oli}. Other constraints have been obtained through the consideration
of the effects of magnetic fields on primordial nucleosynthesis
\cite{kern-sta-vach,grasso-ruben,che oli-sch-tru} and on
the distortion in the microwave background due to the presence of a cosmological
magnetic field \cite{bar-fer-silk,dur}.
Even if a primordial magnetic field was to weak to be of astrophysical
significance, it is still of principal interest to study cosmic magnetic fields
today because they can provide direct and important information about the kind
of physics that must have taken place in the early universe.
There have been quite a few mechanisms proposed for ways of producing magnetic
fields in the early universe. We will not discuss them here but refer to
\cite{hind-ever} for a brief review.
In this work we will not be concerned with any particular model for the
generation of primordial magnetic fields. Instead we will focus on the
universial problem of how a primordial field, whatever its origin, will develop
as the universe evolved. In order to do this one needs to consider
magnetohydrodynamics (MHD) in an expanding universe.
Doing full numerical relativistic MHD simulations of the physics of the early
universe is hard and requires extensive computer memory and time.
Greater dynamical range can be obtained by resorting to approximate methods. The
cascade model of Ref.\ \cite{brand} is one such method, which is thought to
reproduce well the flow of energy between wavenumbers of the full MHD equations,
at the cost of a severe truncation in the number of degrees of freedom. It was
found in \cite{brand} that energy was transferred from small to large scales in
an inverse cascade, and that the correlation length of the initially random
field increased with what looked like a small power of (conformal) time. This
is pleasing if one wants to derive the galactic dynamo seed field from
a primordial process, for general arguments of causality and energy conservation
indicate that such an inverse cascade is actually necessary \cite{hind-ever}.
In this work we will be using a string model approach to relativistic MHD
to study the evolution of cosmic magnetic fields. The connection between MHD
and string dynamics have previously been studied by Semenov \cite{semenov1} and
Olesen \cite{olesen1}. Our approach is simular to that of Semenov \cite{semenov1}
but more general since we do not assume that there is a conserved
particle number density. We take essentially the opposite direction of Olesen \cite{olesen1},
in that we derive string equations from MHD and not the other way around.
Once we have reduced the MHD equations to a string model, the results can be
understood in terms of the coarsening dynamics of cosmic string networks
\cite{vil-shell,hind-kib}. The rate of increase of the network scale length $\xi$
is given by the characteristic velocity of waves on the string, in this
case the Alfv\'en velocity, which decreases as the magnetic field decreases
in strength. The string approach indicates that this decrease in strength
is primarily due to reconnection on small scales: small flux loops are
continually created, transferring energy away from the network of infinitely
long flux lines. The transfer of energy from the large-scale field happens
in a self-similar manner: the magnetic field power spectrum can be displayed
as
\begin{equation}
|{\mathbf{B}}_{{\mathbf{k}}}|^2 \propto \tau^{-\alpha} P(k\xi(\tau)),
\end{equation}
where ${\mathbf{k}}$ is wavenumber and $\tau$ conformal time.
A powerful scaling argument
due to Olesen \cite{olesen2} shows, in the limit of ideal MHD,
that $\xi \propto \tau^{2/(n+5)}$, where
the initial power spectrum behaves as $k^n$ at low $k$.
Causality dictates that
$n \ge 2$ \cite{dur} (and not $n\ge 0$ as one of us \cite{hind-ever} and
another author \cite{son} has stated). As we violate
the ideal condition by allowing reconnection, it is not clear that this
is the correct power of $\tau$.
This scaling law is our main result. We see no sign of a true inverse cascade,
in the sense that power is not transferred from small to large scales. If
anything, the transfer is from large to small, and it is only because energy
is being lost faster from small scales that we see an increase in the scale
length $\xi$.
\section{Relativistic MHD and strings}
In this work we will concentrate on the ideal limit of MHD. This means that we
neglect any viscous effects and treat matter as a perfect fluid. This is a good
approximation at sufficiently large scales. During the radiation
dominated era, which we are mainly concerned with here, the universe was a very
good conductor \cite{turn-widr,aho-enq}. We therefore consider the
$\sigma\rightarrow\infty$ limit of MHD where magnetic diffusion can be ignored
and the magnetic field can be considered to be frozen into the plasma and thus
conserving magnetic flux.
The starting point for ideal relativistic MHD is the energy-momentun tensor
\begin{equation}T^{\mu\nu}=(\rho c^{2}+p)\frac{U^{\mu}U^{\nu}}{c^{2}}-pg^{\mu\nu}+\frac{1}{4\pi}({F^{\mu}}_{\gamma}F^{\nu\gamma}-\frac{1}{4}g^{\mu\nu}F_{\gamma\delta}F^{\gamma\delta}) \label{ideal_e.m.tensor} \end{equation}
consisting of the ideal fluid part and electromagnetic part of the energy-momentum tensor.
Here $p$ is the fluid pressure, $\rho c^{2}$ is the energy density of the fluid, $U^{\mu}$
is the four-velocity of the fluid satisfying the normalisation condition
$U^{\mu}U_{\mu}=c^{2}$ and $F^{\mu\nu}$ is the electromagnetic field tensor.
The evolution equations for the system are given by
\begin{eqnarray}
{T^{\mu\nu}}_{;\nu}&=&0 \label{cov} \\
{F^{\mu\nu}}_{;\nu}&=&\frac{4\pi}{c}J^{\mu} \label{max1} \\
^*\!{{F}^{\mu\nu}}_{;\nu}&=&0 \label{max2} \\
\sigma F^{\mu\nu}U_{\nu}&=&J^{\mu}-J^{\nu}U_{\nu}U^{\mu} \label{ohm}
\end{eqnarray}
where equation (\ref{cov}) expresses covariant energy-momentum conservation,
equations (\ref{max1}) and (\ref{max2}) are Maxwell's equations with $J^{\mu}$ being the
four-current density. In equation (\ref{max2}) $^*\!{{F}^{\mu\nu}}$ is the dual
field tensor defined through the relation
\begin{equation}^*\!{{F}_{\mu\nu}}=\frac{1}{2}\epsilon_{\mu\nu\gamma\rho}F^{\mu\nu} \label{dual} \end{equation}
where $\epsilon_{\mu\nu\gamma\rho}$ is the Levi-Cevita symbol.
Equation (\ref{ohm}) is the relativistic version of Ohm's law where
$\sigma$ is the conductivity of the fluid, measured in the fluid rest frame.
We now repeat the derivation in Subramanian and Barrow \cite{sub-bar} to show that the
evolution equations are conformally invariant and the
evolution can therefore be transformed from the expanding universe to a flat
(Minkowski) spacetime. In so doing we obtain an equivalent set of equations
which are easier to handle.
Two metrics $g_{\mu\nu}$ and $\tilde{g}_{\mu\nu}$ are said to be conformally
related to each other if $\tilde{g}_{\mu\nu}=\Omega^{2}g_{\mu\nu}$ where
$\Omega$ is a non-zero differentiable function.
The flat Robertson-Walker line element has the form
\begin{equation}ds^{2}=dt^{2}-a^{2}(t)d\mathbf{x}^{2} \label{rw} \end{equation}
where $t$ is the comoving proper time and $a(t)$ is the scale factor.
This metric describes a isotropic and homogeneous universe with zero curva\-ture.
The appearence of a hypothetical primordial magnetic field of some strength need
not violate the assumption of isotropy and homogeniety because although the
presence of a magnetic field will locally generate bulk motions in the fluid, if
we look at sufficiently large scales isotropy and homogeniety will be regained.
At large scales the magnetic field, whatever its origin, can be considered as
essentially random since the correlation length of the field is bounded from
above by not exceeding the causal horizon. This justifies the use of the
Robertson-Walker metric.
We introduce conformal time $\tau$ defined by
$d\tau=a^{-1}dt$ so that equation (\ref{rw}) becomes
\begin{equation}ds^{2}=a^{2}(\tau)(d\tau^{2}-d\mathbf{x}^{2}) \end{equation}
and hence $\eta_{\mu\nu}=\tilde{g}_{\mu\nu}=\Omega^{2}g_{\mu\nu}$
with $\Omega=a^{-1}(\tau)$.
We note that under conformal transformations the ideal energy-momentum tensor
$T^{\mu\nu}$ transformation as $T^{\mu\nu}=a^{-6}\tilde{T}^{\mu\nu}$. That this
is so can be seen directly from the definition of the energy-momentum tensor
\begin{equation}T^{\mu\nu}=\frac{2}{\sqrt{-g}}\frac{\partial}{\partial g_{\mu\nu}}(\sqrt{-g}L_{matter}) \end{equation}
The new scaled fields (denoted by tilde) obey ordinary energy-momentum conservation.
To see this we note that the ideal energy-momentum tensor is traceless,
$T\equiv {T^{\mu}}_{\mu}=0$, provided the perfect fluid has the equation of state
$p=\frac{1}{3}\rho c^{2}$. For most of the period before decoupling, the early
universe was radiation dominated and one can use the above equation of state.
We have
\begin{equation}{T^{\mu\nu}}_{;\nu}={T^{\mu\nu}}_{,\nu}+\Gamma^{\mu}_{\nu\rho}T^{\nu\rho}+\Gamma^{\nu}_{\sigma\nu}T^{\sigma\mu} \end{equation}
Using
\begin{equation}\Gamma^{\mu}_{\sigma\rho}=\frac{\dot{a}}{a}(\delta^{\mu}_{\sigma}\delta^{0}_{\rho}+\delta^{\mu}_{\rho}\delta^{0}_{\sigma}-g_{\sigma\rho}\delta^{\mu}_{0})\quad \textrm{and} \quad \Gamma^{\mu}_{\sigma\rho}=4\frac{\dot{a}}{a}(\delta^{\mu}_{\sigma}) \end{equation}
we get
\begin{equation}{T^{\mu\nu}}_{;\nu}={T^{\mu\nu}}_{,\nu}+2\frac{\dot{a}}{a}T^{\mu 0}-\frac{\dot{a}}{a}\delta^{\mu}_{0}T+4\frac{\dot{a}}{a}T^{\mu 0} \end{equation}
But since $T=0$ then
\begin{equation}(a^{-6}\tilde{T}^{\mu 0})_{,\nu}+6\frac{\dot{a}}{a}(a^{-6}\tilde{T}^{\mu 0})=({\tilde {T}}^{\mu\nu}{}_{,\nu})a^{-6}=0 \end{equation}
and hence
\begin{equation}{\tilde {T}}^{\mu\nu}{}_{,\nu}=0 \end{equation}
This means that under conformal transformations our original variables will
transform to a set of new scaled variable satisfying the following relations:
$\rho=a^{-4}\tilde{\rho}$, $p=a^{-4}\tilde{p}$,
$U^{\mu}=a^{-1}\tilde{U}^{\mu}$, $J^{\mu}=a^{-4}\tilde{J}^{\mu}$,
$F^{\mu\nu}=a^{-4}\tilde{F}^{\mu\nu}$.
Now consider Maxwell's equation (\ref{max1}).
Since $F^{\mu\nu}$ is anti symmetric the left hand side simplifies to
\begin{equation}{F^{\mu\nu}}_{;\rho}={F^{\mu\nu}}_{,\rho}+\Gamma^{\nu}_{\sigma\nu}F^{\mu\sigma}=a^{-4}{\tilde {F}}^{\mu\nu}{}_{,\rho} \end{equation}
So the equation for the scaled fields becomes
\begin{equation}{\tilde {F}}^{\mu\nu}{}_{,\nu}=\frac{4\pi}{c}\tilde{J}^{\mu} \end{equation}
For the four-velocity we have
\begin{equation}U_{\nu}=g_{\nu\rho}U^{\rho}=a^{2}\tilde{g}_{\nu\rho}(a^{-1}\tilde{U}^{\rho})=a\tilde{U}_{\nu} \end{equation}
Hence
\begin{equation}\sigma F^{\mu\nu}U_{\nu}=\sigma a^{-3}\tilde{F}^{\mu\nu}\tilde{U}_{\nu}=a^{-4}(\tilde{J}^{\mu}+\tilde{J}^{\nu}\tilde{U}_{\nu}\tilde{U}^{\mu}) \end{equation}
So Ohm's law remains invariant under conformal transformations if we define the
scaled conductivity through $\sigma=\tilde{\sigma}a^{-1}$
So we arrive at the fundamental equations of relativistic MHD,
\begin{eqnarray}
{\tilde {T}}^{\mu\nu}{}_{,\nu}&=&0 \\
{\tilde {F}}^{\mu\nu}{}_{,\nu}&=&\frac{4\pi}{c}\tilde{J}^{\mu} \\
^*\!{\tilde {F}}^{\mu\nu}{}_{,\nu}&=&0 \\
\tilde{\sigma}\tilde{F}^{\mu\nu}U_{\nu}&=&\tilde{J}^{\mu}-\tilde{J}^{\nu}\tilde{U}_{\nu}\tilde{U}^{\mu}/c^2
\end{eqnarray}
From here on we drop the tilde, on the understanding that we mean scaled fields.
We will now introduce a new set of coordinates which will enable us to write the
MHD equations as non-linear string equations.
We define a magnetic four-vector $b^{\mu}$ through the relation
\begin{equation}b^{\mu}=\, ^*\!{F}_{\mu\nu}U^{\nu}/c \end{equation}
We also define new coordinates $x'=(\eta,\sigma,\psi,\zeta)$ such that
$\eta$ are coordinate lines of fluid elements and $\sigma$ are coordinate lines
of magnetic flux.
Hence $U'^{\mu}=(cq,0,0,0)$ and $b'^{\mu}=(0,\beta,0,0)$ satisfying
$U'^{2}=U^{2}=c^{2}$ and $b'^{2}=b^{2}=-B^{2}$ respectively.
Thus we have the metric tensor in the new coordinates
\[g'_{\mu\nu}=diag(1/q^{2},-B^{2}/\beta^{2},-h_{AB}) \]
where $A,B=2,3$.
The new coordinate vectors are
\begin{equation}\frac{\partial x^{\mu}}{\partial\eta}=\frac{U^{\mu}}{q} \end{equation}
\begin{equation}\frac{\partial x^{\mu}}{\partial\sigma}=\frac{b^{\mu}}{\beta} \end{equation}
Since the introduction of these coordinates relies on the frozen in property
of the plasma we will refer to them as frozen-in coordinates.
In the frozen-in coordinate system we can trace the trajectories of fluid
elements by simply varying the value of our time coordinate $\eta$ and keeping
the values of the other coordinated fixed. Similarly, we can trace the magnetic
field lines in the frozen-in system by varying the value of $\sigma$ and keeping
the other three coordinates at fixed values.
The analysis of the MHD equations is usually performed in terms of the magnetic
field and velocity distributions. However, in the description of MHD phenomena,
the concept of a magnetic flux tube is often introduced. The reason is that it
is sometimes convenient, and we gain a better physical insight, if we base the
description on this concept rather on the magnetic field and velocity
distributions. A magnetic flux tube is defined as the volume $V$ enclosed by a
closed surface $\Sigma$ which is everywhere parallel to the ambient magnetic
field vector, and two cross-sectional surface areas $S_{A}$ and $S_{B}$ at
either end. The flux tube therefore consists of a bundle of magnetic field
lines which enter and exit the volume through the end of surfaces $S_{A}$ and
$S_{B}$.
The Cartesian coordinate system, traditionally employed in MHD does not really
lend itself to an analysis of the magnetic flux tube bahaviour. The frozen-in
coordinate system on the other hand, does. The frozen-in coordinates provide a
coordinate system co-moving with a flux tube, and is therefore the more
natural choice for the mathematical analysis of flux tube behaviour.
We now consider the equation for magnetic evolution
\begin{equation}^*\!{F^{\mu\nu}}_{;\nu}=0 \end{equation}
Since $^*\!{F}^{\mu\nu}$ is antisymmetric the divergence is given by
\begin{equation}\frac{1}{\sqrt{-g}}\partial_{\nu}(\sqrt{-g}\,^*\!{F^{\mu\nu}})=0 \end{equation}
Using the fact that we can express the dual field tensor as
\begin{equation}^*\!{F}^{\mu\nu}=(b^{\mu}U^{\nu}-U^{\mu}b^{\nu})/c \label{F_b_U}\end{equation}
we have
\begin{equation}\frac{1}{\sqrt{-g}}\partial_{\nu}\left(\sqrt{-g}q\beta\Big(\frac{b^{\mu}}{\beta}\frac{U^{\nu}}{q}-\frac{U^{\mu}}{q}\frac{b^{\nu}}{\beta}\Big)\right)=0 \end{equation}
which gives
\begin{eqnarray}
\lefteqn{q\beta\left(\Big(\frac{U^{\nu}}{q}\Big)\partial_{\nu}\Big(\frac{b^{\mu}}{\beta}\Big)-\Big(\frac{b^{\nu}}{\beta}\Big)\partial_{\nu}\Big(\frac{U^{\mu}}{q}\Big)\right)+{} }\nonumber\\
& &{}\frac{1}{\sqrt{-g}}\left(\frac{b^{\mu}}{\beta}\partial_{\nu}(\sqrt{-g}\beta U^{\nu})-\frac{U^{\mu}}{q}\partial_{\nu}(\sqrt{-g}q b^{\nu})\right)=0
\end{eqnarray}
But first square bracket is the Lie derivative and so vanishes, giving
\begin{equation}\frac{b^{\mu}}{\beta}\partial_{\nu}(\sqrt{-g}\beta U^{\nu})-\frac{U^{\mu}}{q}\partial_{\nu}(\sqrt{-g}q b^{\nu})=0 \end{equation}
Thus we have
\begin{equation}\partial_{\nu}(\sqrt{-g}\beta U^{\nu})=0 \end{equation}
and
\begin{equation}\partial_{\nu}(\sqrt{-g}q b^{\nu})=0 \end{equation}
Hence in our comoving frame we get
\begin{equation}\frac{\partial}{\partial\eta}(\sqrt{-g}\beta q)=\frac{\partial}{\partial\eta}(B\det(h_{AB}))=0 \end{equation}
\begin{equation}\frac{\partial}{\partial\sigma}(\sqrt{-g}\beta q)=\frac{\partial}{\partial\sigma}(B\det(h_{AB}))=0 \end{equation}
And so we see that
\[B\det(h_{AB})=F(\psi,\zeta) \]
where $F(\psi,\zeta)$ is an arbitrary function of $\psi,\zeta$.
We therefore have the freedom to choose $\det(h_{AB})$ such that $F(\psi,\zeta)=\bar{B}$
where $\bar{B}$ is a constant and so
\[h_{AB}=\frac{\bar{B}}{B}\delta_{AB} \]
which means that we can write
\[\sqrt{-g}=\frac{\bar{B}}{\beta q} \]
We now study the equations of motion, starting from the energy-momentum tensor
given in equation (\ref{ideal_e.m.tensor}). Using the above expression, equation
(\ref{F_b_U}), for the dual field tensor
and the connection between the field tensor and its dual, equation (\ref{dual}),
we can write the energy-momentun tensor in the following form
\begin{equation}T^{\mu\nu}=\left(\rho c^{2}+p+\frac{B^{2}}{4\pi}\right)\frac{U^{\mu}U^{\nu}}{c^{2}}-\frac{b^{\mu}b^{\nu}}{4\pi}-g^{\mu\nu}\left(p+\frac{B^{2}}{8\pi}\right) \end{equation}
From energy-momentum conservation we have
\begin{eqnarray}
{T^{\mu\nu}}_{,\nu}&=&\frac{(\rho c^{2}+p+B^{2}/4\pi))}{\beta}\frac{U^{\mu}}{c}\partial_{\nu}(\beta U^{\nu})
\nonumber\\
&+&\beta(U\cdot\partial)\left(\frac{(\rho c^{2}+p+B^{2}/4\pi)}{\beta}
\frac{U^{\mu}}{c}\right)- \nonumber\\
& &{}\frac{1}{4 \pi q}\partial_{\nu}(q b^{\nu})-\frac{q}{4\pi}
(b\cdot\partial)\left(\frac{b^{\mu}}{q}\right)-g^{\mu\nu}P_{,\nu}=0
\end{eqnarray}
Here $P$ is the total pressure from both fluid and electromagnetic field.
Note that the first and the third terms in this equation vanish.
Writing this equation in our comoving reference frame we find
\begin{equation}
\beta q\left[\frac{\partial}{\partial\eta}
\left(\Big(\rho c^{2}+p+\frac{B^{2}}{4\pi} \Big)\frac{q}{c\beta}
\frac{\partial x^{\mu}}{\partial\eta}\right)-
\frac{1}{4\pi}\frac{\partial}{\partial\sigma}\left(\frac{\beta}{q}\frac{\partial x^{\mu}}
{\partial\sigma}\right)\right]-g^{\mu\nu}P_{,\nu}=0 \label{stringeq}
\end{equation}
Equation (\ref{stringeq}) is the equation of motion in the frozen-in coordinates.
In the frozen-in coordinate system the MHD equation of motion reduces to a set
of non-linear string equations. The behaviour of a magnetic flux tube is
therefore formally analogous to that of a non-linear string.
The last term of the left hand side of equation (\ref{stringeq}) take account of
inhomogeneity (i.e.\ pressure gradients) and it describes the coupling between
neighbouring flux tubes whilst moving through a non-uniform plasma medium.
To summarize, we have shown that the behaviour of a magnetic flux tube is
formally analogous to that of a string and one can therefore model a magnetised
plasma as a fluid composed of strings. We have relied heavily on the
arguments of Semenov and Semenov and Berkinov \cite{semenov1}, with one
improvement: we have dropped their assumption that there is a conserved
particle number density, which is neither necessary nor generally applicable
in the early Universe. Our derivation is also complementary to that of
Olesen \cite{olesen1}, who starts from the relativistic string equations
and shows that they can be interpreted as describing the motion of narrow
flux tubes, providing the total pressure remains constant across the tube.
\section{Approximate string equations}
Exploiting the freedom to change coordinates in the $\sigma,\eta$ subspace,
we choose $\beta$ such that $\beta=qB$ and using this in the equation
of motion (\ref{stringeq}) we have
\begin{equation}
\frac{1}{B}\frac{\partial}{\partial\eta}\left(\frac{c\,B}{v_{A}^{2}}
\frac{\partial x^{\mu}}{\partial\eta}\right)-\frac{1}{B}
\frac{\partial}{\partial\sigma}\left(B\frac{\partial x^{\mu}}
{\partial\sigma}\right)-\frac{4\pi}{B^{2}q^{2}}g^{\mu\nu}P_{,\nu}=0,
\label{stringeq2}
\end{equation}
where we have defined the relativistic Alfv\'en velocity as
\begin{equation}
v_{A}=\frac{c\,B/\sqrt{4\pi}}{\left(\rho c^{2}+p+B^{2}/4\pi\right)^{1/2}}.
\label{alfven}
\end{equation}
Rearranging equation (\ref{stringeq2}) we can write it as
\begin{equation}
\frac{1}{v_{A}}\frac{\partial}{\partial\eta}\left(\frac{1}{v_{A}}\frac{\partial x^{\mu}}{\partial\eta}\right)-
\frac{\partial^{2}x^{\mu}}{\partial\sigma^{2}}=
-\frac{1}{B}\frac{\partial}{\partial\eta}\left(\frac{c\,B}{v_{A}}\right)
\frac{\partial x^{\mu}}{\partial\eta}+
\frac{1}{B}\frac{\partial B}{\partial\sigma}
\frac{\partial x^{\mu}}{\partial\sigma}+
\frac{4\pi}{B^{2}q^{2}}g^{\mu\nu}P_{,\nu}
\label{stringeq3}
\end{equation}
We now argue that for the particular situation we are interested in, it is
justified to neglect the three terms on the right hand side of equation (\ref{stringeq3}).
The third term on the RHS of equation (\ref{stringeq3}) can in general not be
neglected because in many astrophysical situations pressure gradients are
important. However, in the early universe the pressure from the magnetic field should be
much smaller than the fluid radiation pressure and since the early universe had a very
high degree of homogeniety it follows that the gradients of the total pressure
are small.
The second term on the RHS of equation (\ref{stringeq3}) does not have have a
definite sign and so if averaged over time it will be zero. Dropping this term
is equivalent to replacing $B$ by its root mean square value.
Again using the fact that the fluid radiation pressure in the early universe
was much larger than the pressure from the magnetic field and remembering
expression (\ref{alfven}) for the relativistic Alfv\'en velocity it is seen that
the first term on the right side of equation (\ref{stringeq3}) is indeed small
since the ratio $B/v_{A}$ is approximatly constant, thus justifying our decision
to neglect it.
We now rescale the time parameter $\eta$ to $\bar{\eta}$ through the relation
\begin{equation}
\frac{\partial}{\partial\bar{\eta}}=\frac{c}{v_{A}(\eta)}
\frac{\partial}{\partial\eta},
\end{equation}
where we have called attention to the fact that the Alfv\'en velocity may
be time-dependent.
Using the above mentioned approximations and our new time parameter $\bar{\eta}$
we are left with the following equation of motion
\begin{equation}
\frac{1}{c^{2}}\frac{\partial^{2}x^{\mu}}{\partial\bar{\eta}^{2}}-
\frac{\partial^{2}x^{\mu}}{\partial\sigma^{2}}=0. \label{e:str_eqn}
\end{equation}
We also recall that the four-dimensional orthogonality of the coordinates, and
the coordinate choice enforced by $\be = qB$,
supplements this equation with constraints
\begin{equation}
\dot x \cdot x' = 0, \qquad \frac{1}{c^2}\dot x^2 + x^{'2} = 0,
\label{e:str_con}
\end{equation}
where dot (prime)
denotes derivative with respect to the timelike parameter $\bar{\eta}$ $(\sigma)$.
We note that the equation (\ref{e:str_eqn}) and constraint (\ref{e:str_con}) is
of exactly the same form as the equation for a Nambu-Goto string in Minkowski
spacetime expressed in the conformal gauge \cite{vil-shell,hind-kib}. With
ideal Nambu-Goto strings in
Minkowski spacetime one can further take $x^0 = c\bar{\eta}$, to obtain the system
\begin{eqnarray}
\frac{1}{c^{2}}\ddot{\mathbf{x}}-\mathbf{x}''&=&0,\label{codeeq} \\
\dot{\mathbf{x}}\cdot\mathbf{x}' &=& 0 \label{codecon1} \\
\frac{1}{c^{2}}\dot{\mathbf{x}}^{2}+\mathbf{x}^{'2} &=& 1 \label{codecon2}
\end{eqnarray}
This is not possible in general for MHD strings, for one can easily verify that
\begin{eqnarray}
\dot{\mathbf{x}}\cdot\mathbf{x}' &=& c\frac{\gamma^2}{q^2}\frac{{\mathbf{v}} \cdot {\mathbf{B}}}{v_A B},\\
\frac{1}{c^{2}}\dot{\mathbf{x}}^{2}+\mathbf{x}^{'2} &=& \frac{\gamma^2}{q^2}
\left( \frac{{\mathbf{v}}^2}{v_A^2} + \frac{({\mathbf{v}} \cdot {\mathbf{B}})^2}{c^2B^2}\right).
\end{eqnarray}
where $\gamma=(1-\mathbf{v}/c)^{-1/2}$ is the usual relativistic gamma factor.
However, as long as ${{\mathbf{v}} \cdot {\mathbf{B}}} = 0$ in the initial conditions, the
first constraint is preserved by the evolution. Furthermore, we can use our
remaining coordinate freedom to define $q$ by
\begin{equation}
q^2 = \gamma^2 \frac{{\mathbf{v}}^2}{v_A^2},
\label{e:qchoice}
\end{equation}
in which case we really do reproduce the Nambu-Goto equations. We should bear
in mind however that we have made several approximations on the way, which are
worth reiterating.
\begin{enumerate}
\item We have neglected pressure gradients.
\item We have made a kind of mean-field approximation in replacing the magnetic field
by its root mean square value.
\item We have neglected the effect of the time-dependence of $B/v_A$.
\end{enumerate}
We have argued that all this approximations are reasonable in the context
of the early Universe, and find that the Nambu-Goto equations can
be used to approximate a class of MHD velocity and magnetic field configurations
which satisfy ${\mathbf{v}} \cdot {\mathbf{B}} = 0$. They may also be a
reasonable approximation to other configurations,
provided we understand that the constraints are satisfied in an average sense.
\section{Algorithm, simulations and results}
The equation of motion (\ref{codeeq}) can be evolved using the Smith-Vilenkin
algorithm \cite{smith-vil}.
The Smith-Vilenkin algorithm provides an exact discrete evolution for a string network
defined on a face-centered cubic lattice and the evolution equations are
\begin{equation}\mathbf{x}(\eta+\delta,\sigma)=\frac{1}{2}[\mathbf{x}(\eta,\sigma+\delta)+\mathbf{x}(\eta,\sigma-\delta)]+\mathbf{v}(\eta,\sigma)\delta \end{equation}
\begin{equation}\mathbf{v}(\eta+\delta,\sigma)=\frac{1}{2}[\mathbf{v}(\eta,\sigma+\delta)+\mathbf{v}(\eta,\sigma-\delta))]+\frac{1}{4\delta}[\mathbf{x}(\eta,\sigma+2\delta)-2\mathbf{x}(\eta,\sigma)+\mathbf{x}(\eta,\sigma-2\delta)] \end{equation}
By using a leap-frog method, updating the positions and velocities at alternate timesteps,
the above equations allow us to calculate the exact future evolution of
$\mathbf{x}$ and $\mathbf{v}$ from some appropriately choosen initial conditions.
Initial string configurations are generated by a method due to Vachaspati and
Vilenkin \cite{vach-vil}. They considered string formation in a global $U(1)$-
model. The $U(1)$-manifold is discretized by allowing the phase to take only
three possible values. These values are then placed randomly on the sites of the
cubic lattice. As we go around the face of a cube in real space, the phase
descibes a certain trajectory in the group space. A string passes through the
face of a cube if that trajectory has a non-zero winding number. With this method
the string segments join up to form either closed loops or else open strings
which intersect the boundaries of the cubic lattice. The shape of the strings
will be Brownian with step size corresponding to the cell size of the lattice.
We have seen how relativistic MHD under the approximations discussed above can
be described in terms of magnetic flux tubes satisfying the Nambu-Goto equations of
motion. In order to represent the continuous distribution of magnetic flux by
a network of string, we gather together a flux $\Phi$ into ideal Nambu-Goto strings
at positions $X^i(\bar\eta,\sigma)$, with \cite{olesen1}
\begin{equation}
B^{i}(\tau,\mathbf{x})=
\Phi\int d\sigma\frac{\partial X^i(\bar{\eta},\sigma)}{\partial\sigma}
\delta^{3}\left(\mathbf{x}-\mathbf{X}(\bar{\eta},\sigma)\right)
\end{equation}
A few words need to be said about reconnection. In real fluids, magnetic flux
tubes interact and reconnections take place when the field lines cross each
other. In ideal MHD there are no dissipative or viscous effects. Physical
reconnection between field lines cannot take place without resistive effects and
therefore the topology of the magnetic field lines is frozen in the fluid and
does not change with time.
Reconnection, which is a local process, is difficult to describe and the details
are not well understood. The efficiency of reconnection is not known and neither is how
this process depend on the local geometry involved, like the relative
inclination of the flux tubes. To answer this question, one has to go to
numerical solutions of the underlying theory, that is MHD.
In order to allow for reconnection to take place in our simulations we will take
a more simple approach to reconnection. In our model strings are allowed to
reconnect only if they pass through the same lattice point. When two strings
meet, they intercommute with a probability $P$ and in this work we put $P=1$.
Reconnection here take place instantaneously between the discrete evolution
time steps. This is physically reasonable since the reconnection timescale is
small compared with the evolution of the system.
We use a simple estimate for the characteristic
length scale $\xi$ which is defined by
\begin{equation}
\xi^2=V/L,
\end{equation}
where $V$ is the volume of the
lattice and $L$ is the total length of string in the box, excluding small
loops. In the Smith-Vilenkin algorithm it is in fact possible to have a loop
of zero spatial extent, occupying just one lattice point, which travels at the
wave velocity. If such a loop is formed it does not contribute to the magnetic
field, and we do not count it in the calculation of the total length. It is a
well-known feature of Nambu-Goto simulations that nearly all the string ends
up in this kind of loop \cite{vil-shell,hind-kib}. In the MHD context
we should probably not regard these loops as representing magnetic field
lines at all: the energy is probably being dissipated in a very small-scale
reconnection process. In any case, the fact that loops are generically produced
so small underlines the fact that strings are very efficient at transferring
power from large to small scales.
We have studied simulations of the evolution of the magnetic field on lattices
with sizes $(128\delta)^{3}$ and $(256\delta)^{3}$, using a code originally
developed by Sakellariadou \cite{SakVil88}, which implements both the
Vachaspati-Vilenkin algorithm for the initial conditions and the Smith-Vilenkin
alorithm for the evolution. Periodic boundary conditions
were used in all simulations and the evolution time were resticted to less than
half the box size, since after this time causal influences have had time to
propagate around the box.
Having a representation for the magnetic field in real space we use a three
dimensional Fast Fourier Transform algorithm to get a Fourier mode
representation. The power spectrum for the magnetic field
$|\mathbf{B}_{k}|^{2}$ can then be calculated at every time step by averaging
over the amplitudes of all Fourier modes with a wave number between $k$ and
$k+2\pi/\delta$.
Previously it has been shown that networks of cosmic strings, modelled as Nambu-
Goto strings in Minkowski spacetime tend towords a scaling regime
\cite{graham1}. This means that the characteristic length scale of the network
grows as $\sim c\,t$ where $t$ is Minkowski time.
The characteristic length scale for a primordial magnetic field $\xi$ does not
grow with the horizon. Instead we expect the magnetic field to grow as
$\sim c\,\bar{\eta}$.
It is interesting to see if the magnetic field power spectra show scaling
behaviour. In order to investigate this we express the power spectrum in terms
of a scaling function $P$ which is defined through the following expression
\begin{equation}
|B(k,\bar{\eta})|^{2}=V \frac{\Phi^2 P(k\xi)}{\xi}.
\label{e:scaling}
\end{equation}
Here $V$ is the volume of the box and its appearance is just a normalisation
convention. This form for the scaling function ensures that
that the fluctuations obey the scaling law
\begin{equation}
B^2 = \int \frac{d^{3}k}{(2\pi)^3} |B_{k}|^{2}\propto\xi^{-4}\Phi^2
\end{equation}
which is consistent with our picture of a coherent flux $\Phi$ in a region
of size $\xi$.
Figure 1 shows the chacteristic length scale of the magnetic field versus
$c\,\bar{\eta}$ for a typical ensemble.
The measured scaling function $P(k\xi)$ is displayed in Figure 2. The data was
taken from 12 runs on a $(256\delta)^{3}$ lattice, for values of
$c\,\bar{\eta}$ between $40-60$. It is seen that the power spectrum of the magnetic
field does reach a scaling regime. This means that the evolution of the network
will be self-simular with respect to $c\,\bar{\eta}$.
It is clear that the network does exhibit the property of scaling, with
$\xi= x\!_*\cdot c\,\bar{\eta}$. The scaling amplitude $x\!_{*}$ can be obtained by
looking at the ratio $\xi/\bar{\eta}$ towards the end of the simulation and it
is roughly $x\!_{*}\simeq 0.3$ (see \cite{graham1} for a more accurate determination).
Thus we can write
\begin{equation}
\xi=x\!_{*}\cdot c\,\bar{\eta}=x\!_{*}\int v_{A}(\eta)d\eta.
\end{equation}
Unfortunately, we do not yet know how $v_A$ depends on the time parameter
$\eta$ or the conformal time $\tau$. The Alfv\'en velocity $v_A$ depends
on $B$, which in turn depends on $\Phi$ and $\xi$ through (\ref{e:scaling}).
All we can infer from the information at hand is a consistency relation:
if $\xi \propto \eta^r$ and $\Phi \propto \eta^s$, then $3r = s+1$.
The extra information we need comes from the covariance of ideal MHD
under the scale transformation \cite{olesen2}
\begin{equation}
{\mathbf{x}} \to l{\mathbf{x}}, \quad
\tau \to l^{1-h} \tau, \quad
{\mathbf{v}} \to l^h {\mathbf{v}}, \quad
{\mathbf{B}} \to l^h {\mathbf{B}},
\end{equation}
where $h$ is arbitrary. One can show that under this transformation,
\begin{equation}
V^{-1} |\mathbf{B}_{k}|^{2} \to l^{3+h} V^{-1} |\mathbf{B}_{k}|^{2}.
\end{equation}
If we define a function $\Lambda(k,\tau)$ by
\begin{equation}
V^{-1} |\mathbf{B}_{k}|^{2} = \tau^{(3+h)/(1-h)} \Lambda(k,\tau),
\end{equation}
then we see that under the same transformation
\begin{equation}
\Lambda(k,\tau) \to \Lambda(k/l,\tau l^{1-h}) =\Lambda(k,\tau),
\end{equation}
from which we immediately infer that
\begin{equation}
\Lambda(k,\tau) = \Lambda(k\tau^{1/(1-h)}).
\end{equation}
Furthermore, if $|\mathbf{B}_{k}|^{2}$ behaves as $k^n$
as $k\to 0$, we have
\begin{equation}
|\mathbf{B}_{k}|^{2} \propto \tau^{(3+h+n)/(1-h)} k^n
\end{equation}
in that limit. It is often assumed that the large-scale
power is not affected by small-scale processes \cite{olesen2,son},
in which case $h=-n-3$, and we find the scaling laws (derived by
the same authors)
\begin{equation}
\xi \propto \tau^{2/(n+5)}, \qquad \Phi \propto \tau^{(1-n)/(n+5)}.
\end{equation}
In the early universe individual particles move with relativistic velocities.
However, we expect that the bulk velocity of the fluid $v$ to be
non-relativistic. Hence
\begin{equation}
U^0 = qc \frac{\partial \tau}{\partial \eta} \simeq c,
\end{equation}
and given (\ref{e:qchoice}) we see that, on average,
\begin{equation}
\eta = \sqrt{\langle{\dot{\mathbf{x}}^2}\rangle} \tau /v_A,
\end{equation}
where $\langle{\dot{\mathbf{x}}^2}\rangle$ is the mean square string velocity. Simulations
give this to be $0.36$ \cite{graham1}.
\section{Conclusions}
We have seen how the relativistic MHD equations, with a few
reasonable assumptions, may be recast as string-like equations for the
motion of the flux lines. This allows us enormous gains in dynamic
range in the simulation of a random magnetic field in the early Universe,
without being forced to the ideal limit, for we incorporate
diffusivity by allowing reconnections between the strings.
The result is that we can understand the evolution of magnetic fields in
terms of the evolution of a network of strings, and we find that
the power spectrum quickly evolves to a self-similar or scale-invariant
form, with scale length $\xi$ increasing in time. What this power law
is we are unable to say: ideal MHD predicts $\xi \propto \tau^{2/(n+5)}$,
where $n$ is the low $k$ exponent of the power spectrum, but as we have
departed from ideality by allowing reconnection, we cannot make a prediction.
The increase in scale length comes about by the strings straightening at the
Alfv\'en velocity, while forming very small loops which can dissipate
energy quickly. This is the new feature that the string
formulation brings to light: strings transfer energy from large to small
scales in an extremely efficient manner. Thus, although the scale length
increases, it is because power is preferentially lost from small scales.
Whether it is fair to call this an inverse cascade is a matter of terminology.
What is clear is that the dynamics predicted by the string model of MHD
is certainly not of the right kind to produce seeds for the galactic
dynamo from magnetic fields created in the very early Universe.
There are of course many places where this line of argument is vulnerable.
The model makes approximations which we have tried to highlight. Furthermore,
our string simulations use special string configurations to make gains in
computational efficiency: the strings lie on a cubic lattice to start with,
and one may be suspicious that this may introduce some artifice into the
dynamics. However, the propensity of a string network to scale is firmly
believed, so we are confident that the magnetic field power spectrum will
also scale. What is probably not well approximated is the actual form of the
power spectrum, which betrays the particularly string-like feature of a
$k^{-1}$ tail, due to the fact that all the flux is held to be concentrated in
a narrow tube. Furthermore, the string model may be deficient in its
description of helicity,
which is known to be extremely important in the development of true inverse
cascades \cite{pouquet,corn,FieCar98}. The helicity is represented by the
linking number of the strings, but we are not able to incorporate a local
contribution induced, for example, by twisted tubes of flux.
It may well be that we are
missing some very important dynamics here.
We clearly need to check our results against a non-ideal MHD code, to
see if the predicted self-similar dynamics emerges, and also to find
the correct power law for the scale length. This project is currently in hand.
We are extremely grateful to
Mairi Sakellariadou for the use of her Minkowski space string code.
We have also benefited from conversations with Axel Brandenburg,
Carlo Barenghi, Richard Rijnbeek and Vladimir Semenov.
MH is supported by PPARC grant no.\ GR/L56305.
MC is supported by Centrala Studien\"amden (CSN).
|
\section*{Microscopic chaos from Brownian motion?}
In a recent Letter in {\em Nature}, Gaspard et al.~\cite{G} claimed to
present empirical evidence
for microscopic chaos on a molecular scale from an ingenious experiment
using a time series of the positions of a Brownian particle in a liquid.
The Letter was preceded by a lead
article~\cite{DS} emphasising the fundamental nature of the experiment.
In this note we demonstrate that virtually identical results
can be obtained by analysing a corresponding numerical time series of a
particle in a manifestly microscopically nonchaotic system.
As in Ref.~\cite{G} we analyse the position of a single particle
colliding with many others. We use the Ehrenfest wind-tree model~\cite{E}
where the pointlike (``wind'') particle moves in a plane colliding with
randomly placed fixed square scatterers (``trees'', Fig. 1a). We choose
this model because collisions with the flat sides of the squares do not
lead to exponential separation of corresponding points on initially nearby
trajectories. This means there are no positive Lyapunov exponents which
are characteristic of microscopic chaos here. In contrast the Lorentz model
used in~\cite{G} as an example similar to Brownian motion is a wind-tree
model where the squares are replaced by hard (circular) disks (cf.\cite{G},
Fig.~1) and exhibits exponential separation of nearby trajectories, leading
to a positive Lyapunov exponent and hence microscopic chaos.
Nevertheless, we now demonstrate that the nonchaotic Ehrenfest model
reproduces all the results of Ref.~\cite{G}. A particle trajectory
segment shown in Fig.~1b is strikingly similar to that for the Brownian
particle, (cf.\cite{G}, Fig.~2). Our subsequent analysis parallels that of
Ref.~\cite{G}, where more details may be found. Thus the microscopic
chaoticity is determined by estimating the Kolmogorov-Sinai entropy
$h_{KS}$, using the method of Procaccia and others~\cite{GP,CP}
via the information entropy $K(n,\epsilon,\tau)$
obtained from the frequency with which the partical retraces part of
its (previous) trajectory within a distance $\epsilon$ for $n$
measurements spaced at a time interval $\tau$. Since for the systems
considered here $h_{KS}$ equals the sum of the positive Lyapunov exponents,
the determination of a positive $h_{KS}$ would imply microscopic chaos.
As in~\cite{G} we find that $K$ grows linearly with time
(Fig 1c and~\cite{G} Fig.~3), giving a positive (non-zero) bound on $h_{KS}$
(Fig 1d and~\cite{G} Fig.~4). Indeed our Figs.~1b-d for a microscopically
nonchaotic model are virtually identical with the corresponding figures
2-4 of~\cite{G}. Therefore Gaspard et al. did not prove microscopic
chaos for Brownian motion.
The algorithm of~\cite{GP,CP} as applied here cannot determine the
microscopic chaoticity of Brownian motion since the time interval
between measurements, $1/60\;s$ in~\cite{G}, is so much larger than
the microscopic time scale determined by the inverse collision frequency
in a liquid, approximately $10^{-12}\;s$. A decisive determination of
microscopic chaos would involve, it seems at the very least, a time
interval $\tau$ of the same order as characteristic microscopic time scales.
\noindent
\\
{\bf C. P. Dettmann, E. G. D. Cohen}\\
Center for Studies in Physics and Biology,\\
Rockefeller University,\\
New York, NY 10021, USA\\
\\
{\bf H. van Beijeren}\\
Institute for Theoretical Physics,\\
University of Utrecht,\\
3584 CC Utrecht, The Netherlands
|
\section{Introduction}\label{intro}
The Anderson-type metal-insulator transition (MIT) has been the
subject of investigation for decades since Anderson formulated the
problem in 1958 \cite{anderson}. He proposed that increasing the
strength of a random potential in a three-dimensional (3D) lattice may
cause an ``absence of diffusion'' for the electrons. Today, it is
widely accepted that near this exclusively-disorder-induced MIT the
d.\ c.\ conductivity $\sigma$ behaves as $|E-E_c|^\nu$, where $E_c$ is
the critical energy or the mobility edge at which the MIT occurs, and
$\nu$ is a universal critical exponent \cite{kramer}. Numerical
studies based on the Anderson Hamiltonian of localization have
supported this scenario with much evidence
\cite{kramer,schreiber,bulka,hofstetter,slevin}. In measurements of
$\sigma$ near the MIT in semiconductors and amorphous alloys this
behavior was also observed with varying values of $\nu$ ranging from
$0.5$--$1.3$ \cite{nu,lauinger,stupp}. It is currently believed that
these different exponents are caused by interactions in the system
\cite{belitz}. Indeed, an MIT may be induced not only by disorder but
also by interactions such as electron-electron and electron-phonon
interactions, among others \cite{mott2}. Nevertheless, the
experimental confirmation of the critical behavior of $\sigma$ allows
the use of the Anderson model in order to describe the transition
between the insulating and the metallic states in disordered systems.
Besides for the conductivity $\sigma$, experimental investigations can
also be done for thermoelectric transport properties such as the
thermoelectric power $S$ \cite{lauinger,sherwood,lakner}, the thermal
conductivity $K$ and the Lorenz number $L_0$. The behavior of these
quantities at low temperature $T$ in disordered systems close to the
MIT has so far not been satisfactorily explained. In particular, some
authors have argued that $S$ diverges \cite{sherwood,castellani} or
that it remains constant \cite{sivan,enderby} as the MIT is approached
from the metallic side.
In addition, $|S|$ at the MIT has been predicted \cite{enderby} to be
of the order of $\sim 200\,\mu$V/K. On the other hand, measurements of
$S$ close to the MIT conducted on semiconductors for $T\leq1\,$K
\cite{lakner} and on amorphous alloys in the range
$5\,$K$\leq{T}\leq350\,$K \cite{lauinger} yield values of the order of
0.1-1$\,\mu$V/K. They also showed that $S$ can either be negative or
positive depending on the donor concentration in semiconductors or the
chemical composition of the alloy. The large difference between the
theoretical and experimental values is still not resolved.
The objective of this paper is to study the behavior of the
thermoelectric transport properties for the {\em Anderson model}\/ of
localization in disordered systems near the MIT at low $T$. We
clarify the above mentioned difference in the theoretical calculations
for $S$, by showing that the radius of convergence for the Sommerfeld
expansion used in Refs.\ \cite{castellani,sivan} is zero at the MIT.
We show that $S$ is a finite constant at the MIT as argued in Refs.\
\cite{sivan,enderby}. Besides for $S$, we also compute the $T$
dependence for $\sigma$, $K$, and $L_0$. Our approach is neither
restricted to a low- or high-$T$ expansion as in Refs.\
\cite{castellani,sivan}, nor confined to the critical regime as in
Ref.\ \cite{enderby}.
We shall first introduce the model in Sec.\,\ref{Amodel}. Then in
Secs.\,\ref{linear} and \ref{sec-etc} we review the thermoelectric
transport properties in the framework of linear response and the
present formulations in calculating them. In Sec.\,\ref{method} we
shall show how to calculate the $T$ dependence of these properties.
Results of these calculations are then presented in
Sec.\,\ref{result}. Lastly, in Sec.\,\ref{conclude} we discuss the
relevance of our study to the experiments.
\section{The Anderson Model of Localization}\label{Amodel}
The Anderson model \cite{anderson} is described by the Hamiltonian
\begin{equation}
H=\sum_{i}\epsilon_{i}|i\rangle\langle{i}|+\sum_{i\neq{j}}t_{ij}
|i\rangle\langle{j}| \label{hamilton}
\end{equation}
where $\epsilon_{i}$ is the potential energy at the site $i$ of a
regular cubic lattice and is assumed to be randomly distributed in the
range $[-W/2,W/2]$ throughout this work. The hopping parameters
$t_{ij}$ are restricted to nearest neighbors. For this system, at
strong enough disorder and in the absence of a magnetic field, the
one-particle wavefunctions become exponentially localized at $T=0$ and
$\sigma$ vanishes \cite{kramer}. Illustrating this, we refer to
Fig.\,\ref{dosfig}
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{
\includegraphics{density.eps}
}
\caption{
The density of states of a 3D Anderson model, averaged over many
disorder realizations with $W=12$. The solid vertical lines at
$-E_{c}$ and $E_c$ denote the mobility edges.}
\label{dosfig}
\end{figure}
where we show the density of states $\rho(E)$ obtained by
diagonalizing the Hamiltonian (\ref{hamilton}) with the Lanczos method
as in Ref.\,\cite{milde0,milde}. The states in the band tails with energy
$|E|>E_c$ are localized within finite regions of space in the system
at $T=0$ \cite{kramer}. When the Fermi energy $E_F$ is within these
tails at $T=0$ the system is insulating. Otherwise, if
$|E_{F}|<E_{c}$ the system is metallic. The critical behavior of
$\sigma$ is given by
\begin{equation}
\sigma(E)=\left\{ \begin{array}{cc}
\sigma_{0}\left|1-\frac{E}{E_{c}}\right|^{\nu}, & \quad |E| \leq E_{c}, \\
0, & \quad |E|>E_{c},
\end{array}\right. \label{dc_cond}
\end{equation}
where $\sigma_{0}$ is a constant and $\nu$ is the conductivity
exponent \cite{kramer}. Thus, $E_{c}$ is called the mobility edge
since it separates localized from extended states. At the critical
disorder $W_c=16.5$, the mobility edge occurs at $E_c=0$, all states
with $|E|>0$ are localized \cite{schreiber,bulka} and states with
$E=0$ are multifractal \cite{schreiber,milde0}. The value of $\nu$ has
been computed from the non-linear sigma-model \cite{wegner},
transfer-matrix methods \cite{kramer,slevin}, Green functions methods
\cite{kramer}, and energy-level statistics \cite{hofstetter,els}.
Here we have chosen $\nu= 1.3$, which is in agreement with
experimental results in Si:P \cite{stupp} and the numerical data of
Ref.\ \cite{hofstetter}. More recent numerical results
\cite{kramer,slevin}, computed with higher accuracy, suggest that $\nu
= 1.5\pm 0.1$. As we shall show later, this difference only slightly
modifies our results.
We emphasize that the Hamiltonian (\ref{hamilton}) only incorporates
the electronic degrees of freedom of a disordered system and further
excitations such as lattice vibrations are not included.
For comparison with the experimental results, we measure $\sigma$ in
Eq.\,(\ref{dc_cond}) in units of $\mathrm{\Omega}^{-1}\mbox{cm}^{-1}$.
We fix the energy scale by setting $t_{ij} = 1$ eV. Hence the band
width of Fig.\ \ref{dosfig} is comparable to the band width of
amorphous alloys \cite{haussler}. Furthermore, the experimental
investigations of the thermoelectric power $S$ in amorphous alloys
\cite{lauinger} have been done at high electron filling \cite{private}
and thus we will mostly concentrate on the MIT at $E_c$.
\section{Linear Thermoelectric Effects}\label{linear}
\subsection{Definition of the Transport Properties}
Thermoelectric effects in a system are due mainly to the presence of a
temperature gradient $\mathbf{\nabla}T$ and an electric field
$\mathbf{E}$ \cite{ashcroft}. We recall that in the absence of
$\mathbf{\nabla}T$ with $\mathbf{E}\neq{0}$, the electric current
density $\langle\mathbf{j}\rangle$ flowing at a point in a conductor
is directly proportional to $\mathbf{E}$,
\begin{equation}
\langle\mathbf{j}\rangle=\sigma\mathbf{E}\;. \label{sigma1}
\end{equation}
By applying a finite gradient $\mathbf{\nabla}T$ in an open circuit,
electrons, the thermal conductors, would flow towards the low-$T$ end
as shown in Fig.\,\ref{bar}. This causes a build-up of negative
charges at the low-$T$ end and a depletion of negative charges at the
high-$T$ end. Consequently, this sets up an electric field
$\mathbf{E}$ which opposes the thermal flow of electrons. For small
$\mathbf{\nabla}T$, it is given as
\begin{equation}
\mathbf{E}=S\mathbf{\nabla}{T}\,. \label{tp1}
\end{equation}
This equation defines the \textit{thermopower} $S$. In the Sommerfeld
free electron model of metals, $S$ is found to be directly
proportional to $-T$ \cite{ashcroft}. Note that the negative sign is
brought about by the charge of the thermal conductors.
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{
\includegraphics{bar.eps}
}
\caption{
In an open circuit, a temperature gradient $\nabla{T}$ induces an
electric field $\mathbf{E}$ in the opposite direction which opposes
the thermal flow of electrons.}
\label{bar}
\end{figure}
For small $\mathbf{\nabla}T$, the flow of heat in a system is
proportional to $\nabla{T}$. Fourier's Law gives this as
\begin{equation}
\langle{\mathbf{j}_{q}}\rangle=K(-\mathbf{\nabla}T) \label{tc1}
\end{equation}
where $\langle{\mathbf{j}_{q}}\rangle$ is the heat current density and
$K$ is the thermal conductivity \cite{ashcroft}. At low $T$, the
phonon contribution to $\sigma$ and $K$ becomes negligible compared to
the electronic part \cite{ashcroft}. As $T\rightarrow{0}$, $\sigma$
approaches a constant and $K$ becomes linear in $T$. One can then
verify the empirical law of Wiedemann and Franz which says that the
ratio of $K$ and $\sigma$ is directly proportional to $T$
\cite{wiedemann,chester}. The proportionality coefficient is
known as the Lorenz number $L_0$,
\begin{equation}
L_{0}=\frac{e^2}{k_{B}^2}\frac{K}{\sigma{T}}\, \label{lo1}
\end{equation}
where $e$ is the electron charge and $k_{B}$ is the Boltzmann
constant. For metals, it takes the universal value $\pi^{2}/3$
\cite{ashcroft,chester}. Strictly speaking, the law of Wiedemann and
Franz is valid at very low $T$ ($\lesssim{10}\,$K) and at high (room)
$T$. This is because in these regions the electrons are scattered
elastically. At $T\sim10-100\,$K deviations from the law are observed
which imply that $K/\sigma{T}$ depends on $T$.
In summary, Eqs.\,(\ref{sigma1})-(\ref{lo1}) express the
phenomenological description of the transport properties.
\subsection{The Equations of Linear Response}
A more compact and general way of looking at these thermoelectric
``forces'' and effects is as follows: the responses of a system to
$\mathbf{E}$
and $\mathbf{\nabla}T$ up to linear order \cite{callen} are
\begin{equation}
\langle\mathbf{j}\rangle = |e|^{-1}\left(|e|L_{11}
\mathbf{E}-L_{12}T^{-1}\mathbf{\nabla}T\right) \label{ecurrent}
\end{equation}
and
\begin{equation}
\langle\mathbf{j}_{q}\rangle = |e|^{-2}\left(|e|L_{21}
\mathbf{E}-L_{22}T^{-1}\mathbf{\nabla}T\right) \label{hcurrent}.
\end{equation}
The kinetic coefficients $L_{ij}$ are the keys to calculating the
transport properties theoretically. Using Ohm's law (\ref{sigma1}) in
Eq.\,(\ref{ecurrent}), we obtain
\begin{equation}
\sigma = L_{11}\,. \label{sigma2}
\end{equation}
Also from Eq.\,(\ref{ecurrent}), $S$, measured under the condition of
zero electric current, is expressed as
\begin{equation}
S = \frac{L_{12}}{|e|TL_{11}}\;. \label{tp2}
\end{equation}
With the same condition,
Eq.\,(\ref{hcurrent}) yields
\begin{equation}
K=\frac{L_{22}L_{11}-L_{21}L_{12}}{|e|^{2}TL_{11}}\,. \label{tc2}
\end{equation}
From Eq.\ (\ref{lo1}) $L_0$ is given as
\begin{equation}
L_{0}=\frac{L_{22}L_{11}-L_{21}L_{12}}{(k_{B}TL_{11})^2}\,. \label{lo2}
\end{equation}
Therefore, we will be able to determine the transport properties once
we know the coefficients $L_{ij}$. We note that in the absence of a
magnetic field, as considered in this work, the Onsager
relation $L_{21}=L_{12}$ holds \cite{callen}.
Eliminating the kinetic coefficients in Eqs.\,(\ref{ecurrent}) and
(\ref{hcurrent}) in favor of the transport properties, we obtain
\begin{equation}
\langle\mathbf{j}\rangle =
\sigma\mathbf{E}-\sigma{S}\nabla{T}\label{ecurrent2}
\end{equation}
and
\begin{equation}
\frac{\langle\mathbf{j}_{q}\rangle}{T} = S\langle\mathbf{j}\rangle
-\frac{K\nabla{T}}{T}.\label{entropy}
\end{equation}
Here, $\langle\mathbf{j}_{q}\rangle/T$ is simply the entropy current
density \cite{callen}. Hence, the thermopower is just the entropy
transported per Coulomb by the flow of thermal conductors. According
to the third law of thermodynamics, the entropy of a system and, thus,
also $\langle\mathbf{j}_{q}\rangle/T$ will go to zero as
$T\rightarrow{0}$. We can check with Eqs.\,(\ref{ecurrent2}) and
(\ref{entropy}) that this is satisfied by our calculations in the 3D
Anderson model.
\subsection{Application to the Anderson Transition}
In general, the linear response coefficients $L_{ij}$ are obtained
through the Chester-Thellung-Kubo-Greenwood (CTKG) formulation
\cite{chester,kubo}. The kinetic coefficients are expressed as
\begin{equation}
L_{11}= \int_{-\infty}^{\infty} A(E)
\left[ - \frac{\partial f(E,\mu,T)}{\partial E} \right] dE\,, \label{l11}
\end{equation}
\begin{equation}
L_{12}= - \int_{-\infty}^{\infty} A(E)\left[E-\mu(T)\right]
\left[- \frac{\partial f(E,\mu,T)}{\partial E} \right] dE\,, \label{l12}
\end{equation}
and
\begin{equation}
L_{22}= \int_{-\infty}^{\infty} A(E)\left[E-\mu(T)\right]^2
\left[- \frac{\partial f(E,\mu,T)}{\partial E} \right] dE\,, \label{l22}
\end{equation}
where $A(E)$ contains all the system-dependent features, $\mu(T)$ is
the chemical potential and
\begin{equation}
f(E,\mu,T)=1/\left\{1+\exp([E-\mu(T)]/k_{B}T)\right\}
\label{eq-fermi}
\end{equation}
is the Fermi function. The CTKG approach inherently assumes that the
electrons are noninteracting and that they are scattered elastically
by static impurities or by lattice vibrations. A nice feature of this
formulation is that all microscopic details of the system such as the
dependence on the strength of the disorder enter only in $A(E)$. This
function $A(E)$ can be calculated in the context of the
relaxation-time approximation \cite{ashcroft}. However, an exact
evaluation of $L_{ij}$ is difficult, if not impossible, since it
relies on the exact knowledge of the energy and $T$ dependence of the
relaxation time. In most instances, these are not known.
In order to incorporate the Anderson model and the MIT in the CTKG
formulation, a different approach is taken: We have seen
in Eq.\,(\ref{sigma2}) that the d.c.\ conductivity is just $L_{11}$.
Thus, to take into account the MIT in this formulation, we identify
$A(E)$ with $\sigma(E)$ given in Eq.\,(\ref{dc_cond}).
The $L_{ij}$ in Eqs.\,(\ref{l11})-(\ref{l22}) can now be easily
evaluated close to the MIT without any approximation, once the $T$
dependence of the chemical potential $\mu$ is known. Unfortunately,
this is not known for the experimental systems under consideration
\cite{nu,lauinger,stupp,sherwood,lakner}, nor for the 3D Anderson
model. Thus one has to resort to approximate estimations of $\mu$, as
we do next, or to numerical calculations, as we shall do in the next
sections.
\section{Evaluation of the Transport Coefficients}
\label{sec-etc}
\subsection{Sommerfeld expansion in the metallic regime}
\label{sec-sommerfeld}
Circumventing the computation of $\mu(T)$, one can use that
$-\partial{f}/\partial{E}$ is appreciable only in an energy range of
the order of $k_{B}T$ near $\mu\approx{E_F}$. The lowest non-zero $T$
corrections for the $L_{ij}$ are then accessible by the Sommerfeld
expansion \cite{ashcroft}, provided that $A(E)$ is nonsingular and
slowly varying in this region. Hence, in the limit $T\rightarrow{0}$,
the transport properties are \cite{sommer}
\begin{equation}
\sigma=A(E_F)+\frac{\pi^2}{6}(k_{B}T)^{2}\left.\frac{d^{2}A(E)}
{dE^2}\right|_ {E=E_F}\,,\label{sigma3}
\end{equation}
\begin{equation}
S=- \frac{\pi^{2}k_{B}^{2}T}{3|e| A(E_{F})}
\left.\frac{dA(E)}{dE}\right|_{E=E_F}\,, \label{tp3}
\end{equation}
\begin{equation}
K=\frac{\pi^{2}k_{B}^{2}T}{3e^2}\left\{A(E_F)
-\frac{\pi^{2}(k_{B}T)^2}{3A(E_F)}
\left[\frac{dA(E)}{dE}\right]_{E=E_F}^2 \right\}\,,
\label{tc3}
\end{equation}
and consequently
\begin{equation}
L_0=\frac{\pi^{2}}{3}\left\{1-\frac{\pi^{2}(k_{B}T)^2}{3[A(E_F)]^2}
\left[\frac{dA(E)}{dE}\right]_{E=E_F}^2 \right\}. \label{lo3}
\end{equation}
In the derivations of $S$, $K$, and $L_0$, the term of order $T^2$ in
Eq.\ (\ref{sigma3}) has been ignored as is customary. We remark that
the terms of order $T^2$ in Eqs.\,(\ref{tc3}) and (\ref{lo3}) are
usually dropped, too. In this case in the metallic regime, $L_0$
reduces to the universal value $\pi^2/3$ \cite{ashcroft}.
The above approach was adopted in Refs.\,\cite{castellani} and
\cite{sivan} to study thermoelectric transport properties in the
metallic regime close to the MIT. From Eq.\ (\ref{tp3}), the authors
deduce
\begin{equation}
S=-\frac{\nu\pi^{2}k_{B}^{2}T}{3|e|(E_{F}-E_{c})}\,. \label{tpCAS}
\end{equation}
In the metallic regime, this linear $T$ dependence of $S$ agrees with
that of the Sommerfeld model of metals \cite{ashcroft}. However,
setting $A(E) = \sigma(E)$ at the MIT \cite{castellani} in
Eq.\,(\ref{dc_cond}) is in contradiction to the basic assumption of
the Sommerfeld expansion, since it is not smoothly varying at
$E_{F}=E_c$. Thus identifying $A(E) = \sigma(E)$ in Eqs.\ \ref{sigma3}
- \ref{lo3} is only valid in the metallic regime with $k_B T \ll |E_c
- E_F|$.
\subsection{Exact calculation at $\mu(T)= E_c$}
\label{sec-crit}
A different approach taken by Enderby and Barnes is to fix $\mu=-E_c$
at finite $T$ and later take the limit $T\rightarrow{0}$
\cite{enderby}. Thus, again without knowing the explicit $T$
dependence of $\mu$, the coefficients $L_{ij}$ can be evaluated at the
MIT. For the transport properties they obtain,
\begin{equation}
\sigma=\frac{\sigma_{o}\nu (k_{B}T)^{\nu}I_{\nu}}{\left|E_{c}\right|^{\nu}}\,,
\label{sigmaEB}
\end{equation}
\begin{equation}
S = - \frac{k_{B}}{|e|} \frac{\nu+1}{\nu}
\frac{I_{\nu+1}}{I_{\nu}}\,,
\label{tpEB}
\end{equation}
\begin{equation}
K=\frac{\sigma_{o}
(k_{B}T)^{\nu+2}}{e^{2}T\left|E_{c}\right|^{\nu}}\left[(\nu+2)I_{\nu+2}
-\frac{(\nu+1)^2I_{\nu+1}^2}{\nu{I}_{\nu}}\right]\,, \label{kEB}
\end{equation}
and
\begin{equation}
L_{0}=\left[\frac{(\nu+2)I_{\nu+2}}{\nu{I}_{\nu}}-
\frac{(\nu+1)^{2}I_{\nu+1}^2}{(\nu{I}_{\nu})^2}\right].
\label{loEB}
\end{equation}
Here $I_{1}=\ln2$, $I_{\nu}= (1-2^{1-\nu})\Gamma(\nu)\zeta(\nu)$ for
$\mathrm{Re}(\nu)>0,\;\nu\neq1$, with $\Gamma(\nu)$ and $\zeta(\nu)$
the usual gamma and Riemann zeta functions.
We see that at the MIT, $S$ does not diverge nor go to zero but
remains a universal constant. Its value depends only on the
conductivity exponent $\nu$. This is in contrast to the result
(\ref{tpCAS}) of the Sommerfeld expansion.
In addition, we find that $\sigma\propto T^{\nu}$ and $K\propto
T^{\nu+1}$ as $T\rightarrow{0}$. Hence, $\sigma$ and $K/T$ approach
zero in the same way. This signifies that the Wiedemann and Franz law
is also valid at the MIT recovering an earlier result in Ref.\
\cite{strinati} obtained via diagrammatic methods. However, at the
MIT, $L_0$ does not approach $\pi^2/3$ but again depends on $\nu$. We
emphasize that Eqs.\ (\ref{sigmaEB})-(\ref{loEB}) are exact at $T$
values such that $\mu(T)-E_c=0$ \cite{enderby}. Thus the $T$
dependence of $\sigma$, $S$, $K$, and $L_0$ for a given electron
density can only be determined if one knows the corresponding
$\mu(T)$.
\subsection{High-temperature expansion}
In this section, we will study the lowest-order corrections to the
results obtained before with $\mu(T) = E_c$. We do this by expanding
the Fermi function (\ref{eq-fermi}) for $|E_c - \mu(T)| \ll k_B T$.
In addition, we assume $\mu(T) \approx E_F$ for the temperature range
considered. This procedure gives
\begin{equation}
\hspace*{-0.15cm}
\sigma = L_{11} =
\frac{\sigma_{o}\nu (k_{B}T)^{\nu}}{\left|E_{c}\right|^{\nu}}
\left[
I_{\nu}-(\nu-1)I_{\nu-1}\frac{E_c - E_F}{k_B T}
\right]
\label{eq-ht-sigma}.
\end{equation}
For the thermo\-power, the leading-order correction can be obtained
without expanding $f(E,\mu,T)$ in $L_{11}$ and $L_{12}$. This yields a
constant for $S$ at the MIT \cite{sivan}. We obtain
\begin{equation}
S = - \frac{k_{B}}{|e|}
\left[
\frac{\nu+1}{\nu} \frac{I_{\nu+1}}{I_{\nu}}
+ \frac{E_c - E_F}{k_B T}
\right]\,.
\label{eq-ht-tp}
\end{equation}
For $K$ and $L_0$, we again have to use the expansion of $f(E,\mu,T)$
as in (\ref{eq-ht-sigma}) in order to get non-trivial terms. The
resulting expressions are cumbersome and we thus refrain from showing
them here. We remark that the basic ingredients used in the high-$T$
expansion are somewhat contradictory, namely, the expansion is valid
for high $T$ such that $|E_c - E_F| \ll k_B T$, whereas $\mu(T)=E_F$
is true only for $T=0$.
At present, we thus have various methods of circumventing the explicit
computation of $\mu(T)$. However, their ranges of validity are not
overlapping and it is a priori not clear whether the assumptions for
$\mu(T)$ are justified for $S$ or any of the other transport
properties close to the MIT.
In order to clarify the situation, we numerically compute $\mu(T)$ in
the next section and then use the CTKG formulation to compute the
thermal properties without any approximation.
\section{The Numerical Method}\label{method}
In Eqs.\,(\ref{l11})-(\ref{l22}), the explicit $T$ dependence of the
coefficients $L_{ij}$ occurs in $f(E,\mu,T)$ and $\mu(T)$. More
precisely, knowing $\mu(T)$, it is straightforward to evaluate the
$L_{ij}$. We recall that, for any set of noninteracting particles,
the number density of particles $n$ can be determined as
\begin{equation}
n(\mu,T) = \int_{-\infty}^{\infty}dE \rho(E) f(E,\mu,T)\;, \label{numden}
\end{equation}
where $\rho(E)$ is again the density of energy levels (in the unit
volume) as in Fig.\ \ref{dosfig}. Vice versa, if we know $n$ and
$\rho(E)$ we can solve Eq.\ (\ref{numden}) for $\mu(T)$. The density
of states $\rho(E)$ for the 3D Anderson model has been obtained for
different disorder strengths $W$ as outlined in Sec.\ \ref{Amodel}. We
determine $\rho(E)$ with an energy resolution of at least $0.1$ meV
($\sim 1$ K).
Using $\rho(E)$, we first numerically calculate $n$ at $T=0$ for the
metallic, critical and insulating regimes using the respective Fermi
energies $|E_F| < E_c$, $E_F = E_c$, and $|E_F| > E_c$. With
$\mu=E_{F}$, we have
\begin{equation}
n(E_{F})=\int_{-\infty}^{E_F}dE\rho(E)\,.\label{numEF}
\end{equation}
Next, keeping $n$ fixed at $n(E_F)$, we numerically determine $\mu(T)$
for small $T >0$ such that $|n(E_F) - n(\mu, T)|$ is zero. Then we
increase $T$ and record the respective changes in $\mu(T)$. Using
this result in Eqs.\ (\ref{l11})--(\ref{l22}) in the CTKG formulation,
we compute $L_{ij}$ by numerical integration and subsequently
determine the $T$ dependent transport properties
(\ref{sigma2})--(\ref{lo2}).
We consider the disorders $W=8$, $12$, and $14$ where we do not have
large fluctuations in the density of states. These values are not too
close to the critical disorder $W_c$, so that we could clearly observe
the MIT of Eq.\ (\ref{dc_cond}). The respective values of $E_c$ have
been calculated previously \cite{schreiber} to be close to $7.0$,
$7.5$, and $8.0$. Within our approach, we choose $E_c$ to be equal to
these values.
\section{Results and Discussions}
\label{result}
Here we show the results obtained for $W=12$ with $E_c=7.5$. The
results for $\sigma$, $K$, and $L_0$ are the same at $-E_c$ and $E_c$
since they are functions of $L_{11}$, $L_{22}$ and $L_{12}^{2}$, only.
On the other hand, this is not true for $S$.
\subsection{The Chemical Potential}
In Fig.\,\ref{mufig}, we show how $\mu(T)$ behaves for the 3D Anderson
model at $E_F - E_c = 0$, and $\pm 0.01$. To compare results from
different energy regions we plot the difference of $\mu(T)$ from
$E_F$. We find that $\mu(T)$ behaves similarly in the metallic and
insulating regions and at the MIT for both mobility edges at low $T$.
In all cases we observe $\mu(T)\propto{T}^2$. Furthermore, we see that
$\mu(T)$ at $-E_{c}$ equals $-\mu(T)$ at $E_c$. This symmetric
behavior with respect to $E_{F}=\mu$ reflects the symmetry of the
density of states at $E=0$ as shown in Fig.\,\ref{dosfig}.
For comparison and as a check to our numerics, we also compute with
our method $\mu(T)$ of a free electron gas. The density of states is
\cite{ashcroft}
\begin{equation}
\rho(E)= \frac{3}{2} \frac{n}{E_F}\left( \frac{E}{E_F} \right)^{1/2}
\end{equation}
and we again use $E_F=E_c = 7.5$. We remark that this value of the
mobility edge is in a region where $\rho(E)$ increases with $E$ in an
analogous way as $\rho(E)$ for the Anderson model at $-E_c$ . Thus, as
shown in Fig.\ \ref{mufig}, $\mu(T)$ of a free electron gas is concave
upwards as in the case of the Anderson model at $-E_c$. We also plot
the result for $\mu(T)$ obtained by the usual Sommerfeld expansion for
Eq.\,(\ref{numden}),
\begin{equation}
E_{F}-\mu(T)=\frac{E_F}{3}\left(\frac{\pi k_B T}{2E_F}\right)^2.
\label{fegmu}
\end{equation}
We see that our numerical approach is in perfect agreement with the
free electron result.
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{%
\includegraphics{mu_w12.eps}
}
\caption{
The temperature dependence of the chemical potential $\mu$ measured
with respect to the Fermi energy near both mobility edges. Also
shown is $\mu(T)$ for a free electron gas. The solid line denotes
$\mu(T)$ of Eq.\ (\protect\ref{fegmu}).}
\label{mufig}
\end{figure}
\subsection{The d.c.\ Conductivity}
\label{dcc}
In Fig.\,\ref{dccondfig} we show the $T$ dependence of $\sigma$. The
values of $E_F$ we consider and the corresponding fillings $n$ are
given in Tab.\ \ref{tab:sym}.
\begin{table}[b]
\caption{
Differences of $E_F$ and $n(E_F)$ with respect to the mobility edge
at $E_c=7.5$. The density at $E_c$ corresponds to $n= 97.768\%$.}
\label{tab:sym}
\begin{center}
\begin{tabular}{lr@{}lr@{}lc}
\hline\noalign{\smallskip}
regime &
\multicolumn{2}{c}{$E_F-E_c$} &
\multicolumn{2}{c}{$n(E_F)-n(E_c)$} &
symbol \\
&
\multicolumn{2}{c}{(eV)} &
\multicolumn{2}{c}{($\%$)} &
\\
\noalign{\smallskip}\hline\noalign{\smallskip}
metallic & -0. &010 & -0. &031 & $\circ$ \\
& -0. &007 & -0. &022 & $\bigtriangledown$ \\
& -0. &005 & -0. &015 & $\Box$ \\
& -0. &003 & -0. &009 & $\bigtriangleup$ \\
& -0. &001 & -0. &003 & $\Diamond$ \\
\noalign{\smallskip}
critical & 0. &000 & 0. &000 & $\bullet$ \\
\noalign{\smallskip}
insulating & 0. &001 & 0. &003 & $+$ \\
& 0. &003 & 0. &009 & $\times$ \\
& 0. &010 & 0. &031 & $*$ \\
\noalign{\smallskip}\hline
\end{tabular}
\end{center}
\end{table}
The conductivity at $T=0$ remains finite in the metallic regime with
$\sigma/\sigma_{o}= {|1 - E_F/E_c|}^{\nu}$, because $(-\partial
f/\partial E) \rightarrow \delta(E-E_F)$ in Eq.\ (\ref{l11}) as
$T\rightarrow{0}$. Correspondingly, we find $\sigma=0$ in the
insulating regime at $T=0$.
In the critical regime, $\sigma(T\rightarrow{0})\sim {T}^{\nu}$, as
derived in Ref.\ \cite{enderby}, see Eq.\,(\ref{sigmaEB}). We note
that as one moves away from the critical regime towards the metallic
regime one finds within the accuracy of our data that
$\sigma\sim{T}^2$. We observe that in the metallic regime $\sigma$
increases for increasing $T$. This is different from the behavior in
a real metal where $\sigma$ decreases with increasing $T$. However,
as explained in Sec.\ \ref{Amodel}, the behavior of $\sigma$ in
Fig.\,\ref{dccondfig} is due to the absence of phonons in the present
model.
We also show in Fig.\,\ref{dccondfig} results of the Sommerfeld
expansion (\ref{sigma3}) and the high-$T$ expansion
(\ref{eq-ht-sigma}) for $\sigma$. Paradigmatic for what is to follow
we see that the radius of convergence of the Sommerfeld expansion
decreases for $E_F\rightarrow{E_c}$ and in fact is zero in the
critical regime. On the other hand, the high-$T$ expansion is very
good in the critical regime down to $T=0$ at $E_c=E_F$. The small
systematic differences between our numerical results and the high-$T$
expansion for large $T$ are due to the differences in $\mu(T)$ and
$E_F$. The expansion becomes worse both in the metallic and insulating
regimes for larger $T$. All of this is in complete agreement with the
discussion of the expansions in Sec.~\ref{sec-etc}.
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{
\includegraphics{dccondSF.eps}}
\caption{
The low temperature behavior of the d.c. conductivity $\sigma$. The
symbols are as shown in Tab.\ \protect\ref{tab:sym}.
The dashed lines represent the Sommerfeld expansion result for
$\sigma(T)$ as given in Eq.\,(\protect\ref{sigma3}). For all $8$
choices of $E_F-E_c$, the corresponding high-$T$ expansion
(\protect\ref{eq-ht-sigma}) is indicated by solid lines.}
\label{dccondfig}
\end{figure}
\subsection{The Thermopower}
\label{tp}
In Fig.\,\ref{tpfig1}, we show the behavior of the thermopower at low
$T$ near the MIT. In the metallic regime, we find $S\rightarrow{0}$ as
$T\rightarrow{0}$. At very low $T$, $S\propto T$ as predicted by the
Sommerfeld expansion (\ref{tpCAS}).
We see that the Sommerfeld expansion is valid for not too large values
of $T$. But upon approaching the critical regime, the expansion
becomes unreliable similar to the case of the d.c.\ conductivity of
Sec.\ \ref{dcc}. This behavior persists even if we include higher
order terms in the derivation of $S$ such as the ${\rm O}(T^2)$ term
of Eq.\ (\protect\ref{sigma3}) as shown in Fig.\,\ref{tpfig1}.
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{
\includegraphics{thermopower_BW_W12.eps}}
\caption{
The low temperature behavior of the thermopower $S$. The symbols are
as shown in Tab.\ \protect\ref{tab:sym}.
The dashed lines represent the behavior of $S(T)$ in the metallic
regime as given in Eq.\,(\protect\ref{tpCAS}). The dot-dashed lines
indicate $S$, calculated with the ${\rm O}(T^2)$ term of Eq.\
(\protect\ref{sigma3}), for $E_F-E_c= -0.01$ eV $(\circ)$ and
$-0.001$ eV $(\diamond)$. Solid lines are obtained from the high-$T$
expansion (\protect\ref{eq-ht-tp}). The inset
shows the behavior at $E_F=E_c$ on an enlarged scale. }
\label{tpfig1}
\end{figure}
Before discussing the critical regime in detail, let us turn our
attention to the insulating regime. Here, $S$ becomes very large as
$T\rightarrow{0}$. We have observed that it even appears to approach
infinity. A seemingly divergent behavior in the insulating regime has
also been observed for Si:P \cite{liu}, where it has been attributed
to the thermal activation of charge carriers from $E_F$ to the
mobility edge $E_c$. However, there is a simpler way of looking at
this phenomenon. We refer again to the open circuit in
Fig.\,\ref{bar}. Suppose we adjust $T$ at the cooler end such that
$\nabla{T}$ remains constant. As $T\rightarrow{0}$ both $\sigma$ and
$K$ vanish in the case of insulators --- for $K$ we show this in the
next section. This implies that as $T$ decreases it becomes
increasingly difficult to move a charge from $T$ to
${T}+\delta{T}$. We would need to exert a larger amount of force, and
hence, a larger $\mathbf{E}$ to do the job. From Eq.\,(\ref{tp1}),
this implies a larger $S$ value.
In the critical regime, i.e., setting $E_F = E_c$, we observe in
Fig.\,\ref{tpfig1} that for $T\rightarrow 0$ the thermopower $S$
approaches a value of $228.4\,\mu$V/K. This is exactly the magnitude
predicted \cite{enderby} by Eq.\,(\ref{tpEB}) for $\nu=1.3$. In the
inset of Fig.\,\ref{tpfig1}, we show that the $T$ dependence of $S$ is
linear. The nondivergent behavior of $S$ clearly separates the
metallic from the insulating regime. Furthermore, just as for
$\sigma$, the Sommerfeld expansion for $S$ breaks down at
$E_{F}={E_c}$, i.e., the radius of convergence is zero. Thus, the
divergence of Eq.\,(\ref{tpCAS}) at $E_{F}={E_c}$ reflects this
breakdown and is not physically relevant. On the other hand, the
high-$T$ expansion \cite{sivan} nicely reflects the behavior of $S$
close to the critical regime as also shown in Fig.\ \ref{tpfig1}. For
$E_F = E_c$, the high-$T$ expansion (\ref{eq-ht-tp}) assumes a
constant value of $S$ for all $T$ due to setting $\mu(T)=E_F$. This is
approximately valid, the differences are fairly small as shown in the
inset of Fig.\,\ref{tpfig1}.
We stress that there is no contradiction that $S>0$ in our
calculations whereas $S<0$ in Ref.\,\cite{enderby}. In
Fig.\,\ref{pnfig}, we compare $S$ in energy regions close to $E_{c}$
and to $-E_{c}$ \cite{villa}. Clearly, they have the same magnitude
but $S<0$ at $-E_c$ and $S>0$ at $E_c$. The two cases mainly differ in
their number density $n$. At $-E_c$ the system is at low filling with
$n=2.26\%$ while at $E_c$ the system is at high filling with
$n=97.74\%$. The sign of $S$ implies that at low filling the
thermoelectric conduction is due to electrons and we obtain the usual
picture as in Fig.\,\ref{bar} where the induced field $\mathbf{E}$ is
in the direction opposite to that of $\nabla{T}$. At high filling,
$S>0$ means that $\mathbf{E}$ is directed parallel to $\nabla{T}$.
This can be interpreted as a change in charge transport from electrons
to holes. We remark that this sign reversal also occurs in the
insulating as well as in the critical regime.
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{
\includegraphics{pos_neg.eps}}
\caption{
An example that the magnitude of $S(T)$ is the same in metallic
regions close to $-E_c$ ($\blacksquare$) and $E_c$ ($\circ$). The
$+$-symbols indicate $|S|$ for $-E_c$ and $|E_F-E_c|= 0.01$ eV in all
cases.}
\label{pnfig}
\end{figure}
In Fig.\,\ref{tpfig2}, we take the data of Fig.\,\ref{tpfig1} and plot
them as a function of $\mu-E_c$. Our data coincides with the
isothermal lines which were calculated according to Ref.\
\cite{enderby} by numerically integrating $L_{12}$ and $L_{11}$ for a
particular $T$ to get $S$. We observe that all isotherms of the
insulating ($\mu > E_c$) and the metallic ($\mu < E_c$) regimes cross
at $\mu = E_c$ and $S=228.4\,\mu$V/K. Comparing with
Eq.\,(\ref{tpCAS}), we again find that the Sommerfeld expansion does
not give the correct behavior of $S$ in the critical regime.
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{
\includegraphics{muthermo.eps}}
\caption{
The data of $S$ in Fig.\,\protect\ref{tpfig1} shown as a function of
$\mu$ measured from $E_c = 7.5$ eV. The horizontal line
indicates the fixed point MIT value as given in
Eq.\,(\protect\ref{tpEB}). The thin dashed lines represent
isotherms of $S$ calculated using the same method as in
Ref.\,\protect\cite{enderby}. The solid line is an isotherm of $S$
obtained from Eq.\,(\protect\ref{tpCAS}) for $T=22.3$ K. }
\label{tpfig2}
\end{figure}
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{%
\includegraphics{muscale2.eps}
}
\caption{Scaling plot of the thermopower $S$. The thick dashed line
indicates the fixed point value at the MIT, the solid line
represents the high-$T$ expansion (\protect\ref{eq-ht-tp}),
and the thin dashed line shows the Sommerfeld expansion. The inset
shows the difference in the scaling when plotting $S$ for $E_F-E_c=
-0.001$ eV as function of $(\mu - E_c)/k_B T$ (open
symbols) or $(E_F - E_c)/k_B T$ (filled symbols). }
\label{tpfig3}
\end{figure}
The data presented in Fig.\,\ref{tpfig2} suggest that one can scale
them onto a single scaling curve. In Fig.\,\ref{tpfig3}, we show that
this is indeed true, when plotting $S$ as a function of $(\mu-E_c)/k_B
T$. We emphasize that the scaling is very good and the small width of
the scaling curve is only due to the size of the symbols. The result for
the high-$T$ expansion is indicated in Fig.\,\ref{tpfig3} by a solid
line. It is good close to the MIT. In the metallic regime, the
Sommerfeld expansion correctly captures the decrease of $S$ for large
negative values of $(\mu -E_c)/k_B T$. We remark that a scaling with
$(E_F-E_c)/k_B T$ as predicted in Ref.\ \cite{sivan} is approximately
valid. The differences are very small as shown in the inset of
Fig.\,\ref{tpfig3}.
\subsection{The Thermal Conductivity and the Lorenz Number}
In Fig.\,\ref{thcondfig}, we show the $T$ dependence of the thermal
conductivity $K$. We see that $K\rightarrow{0}$ as $T\rightarrow{0}$
whether it be in the metallic or insulating regime. We note again
that this simple behavior is due to the fact that our model does not
incorporate phonon contributions. The $T$ dependence of $K$ varies
whether one is in the metallic regime or in the insulating regime and
how far one is from the MIT. Directly at the MIT, we find that
$K\rightarrow{0}$ as $T^{\nu+1}$ confirming the $T$ dependence of $K$
as given in Eq.\,(\ref{kEB}). Near the localization MIT, the $T$
dependence of $K/T$ is thus the same as for $\sigma$ in agreement with
Ref.\,\cite{strinati}. Again, we see that the Sommerfeld expansion
(\ref{tc3}) is reasonable only at low $T$ in the metallic regime. As
for $\sigma$ and $S$, we see that the high-$T$ expansion is
again fairly good in the vicinity of the critical regime.
\begin{figure}
\hspace*{0.75cm}
\resizebox{0.4\textwidth}{!}{
\includegraphics{thcond.eps}}
\caption{
The thermal conductivity $K$ as a function of temperature. The
symbols are as shown in Tab.\ \protect\ref{tab:sym}.
The dashed lines were obtained in ${\rm O}(T)$ from the Sommerfeld
expansion (\ref{tc3}) for the metallic regime. The results of the
high-$T$ expansion for the $8$ choices of $E_F-E_c$ are
indicated by solid lines.}
\label{thcondfig}
\end{figure}
At this point we are able to determine the behavior of the entropy in
the system as $T\rightarrow{0}$. In the metallic regime, $S$ and $K$
vanish as $T\rightarrow{0}$, while in the critical and insulating
regime, $\sigma$ and $K$ vanish as $T\rightarrow{0}$. Applying these
results to Eqs.\,(\ref{ecurrent2}) and (\ref{entropy}) yields that for
all regimes the entropy current density $\langle\mathbf{j_q}\rangle/T$
vanishes as $T\rightarrow{0}$. Therefore, we find that the third law
of thermodynamics is satisfied for our numerical results of the 3D
Anderson model.
Next, we present the Lorenz number (\ref{lo1}) as a function of $T$ in
Fig.\,\ref{Lonumfig}. In the metallic regime, we obtain the universal
value $\pi^2/3$ as $T\rightarrow{0}$. Note that for a metal this
value should hold up to room $T$ \cite{ashcroft}. However,
our results for the Anderson model show a nontrivial $T$ dependence.
One might have hoped that the higher-order terms in Eq.\,(\ref{lo3})
could adequately reflect the $T$ dependence of our $L_0$ data.
However, this is not the case as shown in Fig.\,\ref{Lonumfig}. This
indicates that even if we incorporate higher order $T$ corrections the
Sommerfeld expansion will not give the right behavior of $L_0$ near
the MIT. We emphasize that the radius of convergence of
Eq.\,(\ref{lo3}) is even smaller than for $\sigma$, $S$ and $K$.
Similarly, the high-$T$ expansion is also much worse than previously
for $\sigma$, $S$ and $K$. Thus in addition to the results for the
critical regime, we only show in Fig.\,\ref{Lonumfig} the results for
nearby data sets in the insulating and metallic regimes.
The $T$ dependence of $L_0$ is linear as shown in the inset of
Fig.\,\ref{Lonumfig}. As before for $S$, the high-$T$
expansion does not reproduce this. At the MIT, $L_{0}=2.4142$. This
is again the predicted \cite{enderby} $\nu$-dependent value as given
in Eq.\,(\ref{loEB}).
\begin{figure}
\hspace*{0.75cm} \resizebox{0.4\textwidth}{!}{
\includegraphics{LorenzwSF.eps}}
\caption{
The Lorenz number $L_0$ as function of temperature. The
symbols are as shown in Tab.\ \protect\ref{tab:sym}.
The dashed circles mark the values of $L_0$ at $T=0$ for metallic
and insulating regimes. The dashed lines were obtained from Eq.\
(\protect\ref{lo3}). The results of the high-$T$ expansion for
$E_F-E_c= 0$ eV, $\pm 0.001$ eV and
$0.003$ eV
are indicated by solid lines. The inset shows the
behavior at $E_F=E_c$ on an enlarged scale.}
\label{Lonumfig}
\end{figure}
In the insulating regime, one can show analytically by taking the
appropriate limits that $L_0$ approaches $\nu +1$ as $T\rightarrow 0$.
In agreement with this, we find that $L_0=2.3$ at $T=0$ in Fig.\
\ref{Lonumfig}. At first glance, it may appear surprising that a
transport property in the insulating regime could be determined by a
universal constant of the critical regime such as $\nu$. However, in
the evaluation of the coefficients $L_{ij}$, the derivative of the
Fermi function for any finite $T$ decays exponentially and thus one
will always have a non-zero overlap with the critical regime. In the
evaluation of Eq.\ (\ref{lo2}), this $\nu$ dependence survives in the
limit $T\rightarrow 0$. In real materials, we expect the relevant
high-energy transfer processes to be dominated by other scattering
events and thus $L_0$ should be different. Nevertheless, for the
present model, this $\nu$ dependence holds.
\subsection{Possible Scenarios in the Critical Regime}
The results presented in Sec.\ \ref{tp} for the thermopower at the MIT
show that $S=228.4\, \mu$V/K for $\nu=1.3$. This value is $2$ orders of
magnitude larger than those measured near the MIT
\cite{lauinger,sherwood,lakner}. However, as mentioned in the
introduction, the conductivity exponents found in many experiments are
either close $\nu=0.5$ or to $1$ \cite{nu} and one might hope that
this difference may explain the small experimental value of $S$.
Also, recent numerical studies of the MIT by transfer-matrix methods
together with non-linear finite-size scaling find $\nu=1.57\pm{0.03}$
\cite{slevin}. In Tab.\ \ref{tab:nu} we summarize the values of $S$
and $L_0$ at the MIT for these conductivity exponents. We see that
all $S$ values still differ by $2$ orders of magnitude from the
experimental results. Furthermore, we note that our results for $S$
and $L_0$ are independent of the unit of energy. Even if, instead of
$1\,$ eV, we had used $t_{ij}=1\,$ meV, which is appropriate in the doped
semiconductors \cite{nu,stupp,lakner,liu}, we would still obtain the
values as in Tab.\ \ref{tab:nu}.
\begin{table}[b]
\caption{
The thermopower and the Lorenz number at the MIT for a 3D Anderson
model evaluated for various $\nu$ at $E_c=7.5$ eV. The values for
$\nu=0.5$ and $1$ have already been shown in Ref.\
\protect\cite{enderby}.}
\label{tab:nu}
\begin{center}
\begin{tabular}{r@{}lr@{}lr@{}l}
\hline\noalign{\smallskip}
\multicolumn{2}{c}{$\nu$} &
\multicolumn{2}{c}{$S$} &
\multicolumn{2}{c}{$L_0$} \\
& &
\multicolumn{2}{c}{($\mu$V/K) } &
& \\
\noalign{\smallskip}\hline\noalign{\smallskip}
0. & 5 & 163. & 5 & 1. & 7761 \\
1. & 0 & 204. & 5 & 2. & 1721 \\
1. & 3 & 228. & 4 & 2. & 4142 \\
1. & 57 & 249. & 7 & 2. & 6372 \\
\noalign{\smallskip}\hline
\end{tabular}
\end{center}
\end{table}
Thus our numerical results for the thermopower of the Anderson model
at the MIT show a large discrepancy from experimental results. This
may be due to our assumption of the validity of Eq.\ (\ref{dc_cond})
for a large range of energies, or due to the absence of a true
Anderson-type MIT in real materials, or due to problems in the
experiments.
A different scenario for a disorder driven MIT has been proposed by
Mott, who argued that the MIT from the metallic state to the
insulating state is discontinuous \cite{mott1}. Results supporting
such a behavior have been found experimentally \cite{mott2,moebius}.
According to this scenario, $\sigma$ drops from a finite value
$\sigma_{min}$ to zero \cite{mott1} for $T=0$ at the MIT. This minimum
metallic conductivity $\sigma_{min}$ was estimated by Mott to be
\begin{equation}
\sigma_{min}\simeq\frac{1}{a}\frac{e^2}{\hbar} \label{dc_min}
\end{equation}
where $a$ is some microscopic length of the system such as the inverse
of the Fermi wave number, $a\approx{k}_{F}^{-1}$. As summarized in
Ref.\,\cite{mott2}, experiments in non-crystalline materials seem to
indicate that $\sigma_{min}>{300}\;\mathrm{\Omega}^{-1}$cm$^{-1}$.
Let us assume the behavior of $\sigma(E)$ close to the MIT to be
\begin{equation}
\sigma(E)=
\left\{ \begin{array}{cc}
\sigma_{min}, & \quad |E| \leq E_{c}, \\
0, & \quad |E|>E_{c},
\end{array}\right.
\label{mmc}
\end{equation}
with $\sigma_{min}={300}\;\mathrm{\Omega}^{-1}$cm$^{-1}$.
Using the numerical approach of Sec.\ \ref{method}, we obtain
$S=119.5\;\mu$V/K at the MIT. This value is still rather large and
thus the assumption of a minimum metallic conductivity as in Eq.\
(\ref{mmc}) cannot explain the discrepancy from the experimental
results. We remark that the order of magnitude of $S$ is not changed
appreciably, even if we add to the metallic side of Eq.\,(\ref{mmc}) a
term as given in Eq.\,(\ref{dc_cond}) with $\sigma_0$ a few hundred
$\mathrm{\Omega}^{-1}$cm$^{-1}$ and $\nu=1$.
Lastly, we note that the transport properties calculated for $W = 8$
and $14$ do not differ from those obtained for $W = 12$ in both the
metallic and insulating regions provided we are at temperatures $T
\lesssim 100$K. For $S$ and $L_0$ at the MIT we obtain the same
values as for $W = 12$. Again we observe that both $S$ and $L_0$
approach these values linearly with $T$, but with different slopes.
Our results show that the higher the disorder strength the smaller the
magnitude of the slope.
\section{Conclusions}
\label{conclude}
In this paper, we investigated the thermoelectric effects in the 3D
Anderson model near the MIT. The $T$ dependence of the transport
properties is determined by $\mu(T)$. We were able to compute $\mu(T)$
by numerically inverting the formula for the number density $n(\mu,T)$
of noninteracting particles. Using the result for $\mu(T)$, we
calculated the thermoelectric transport properties within the
Chester-Thellung-Kubo-Greenwood formulation of linear response. As
$T\rightarrow{0}$ in the metallic regime we verified that $\sigma$
remains finite, $S\rightarrow{0}$, $K\rightarrow{0}$ and
$L_{0}\rightarrow\pi^{2}/3$. On the other hand, in the insulating
regime, $S\rightarrow\infty$. This we attribute to both $\sigma$ and
$K$ going to zero. Thus, it becomes increasingly difficult to achieve
equilibrium and, hence, the system requires
$\mathbf{E}\rightarrow\infty$. For $L_0$, we obtained a universal
value of $\nu+1$ even in the insulating regime.
Directly at the MIT, the thermoelectric transport properties agree
with those obtained in Ref.\ \cite{enderby}. Namely, as
$T\rightarrow{0}$, we found $\sigma\sim T^{\nu}$, $K\sim T^{\nu+1}$,
while $L_0\rightarrow{\mbox{const}}$.
The thermopower $S$ also remains nearly constant in the critical
regime and, in particular, it does not diverge at the MIT in contrast
to earlier calculations using the Sommerfeld expansion at low $T$
\cite{castellani}. Here we showed that the difference is not so much
due to an order of limits problem, but rather reflects the breakdown
of convergence of the Sommerfeld expansion at the MIT \cite{sivan}.
Our result is supported by scaling data for $S$ at different values of
$T$ and $E_F$ onto a single curve which is continuous across the
transition. Some of the experiments \cite{lauinger,sherwood} for $S$
have been influenced by the Sommerfeld expansion such that the authors
plot their results as $S/T$. We remark that in such a plot the
signature of the MIT is hard to identify, since $S/T$ at the MIT
diverges as $T\rightarrow 0$ solely due to the decrease in $T$. Our
results suggest that plots as in Figs.\,\ref{tpfig1} and \ref{tpfig2}
should show the MIT more clearly.
The value of $S$ is at least two orders of magnitude larger than
observed in experiments \cite{lauinger,sherwood,lakner}. This large
discrepancy may be due to the ingredients of our study, namely, we
assumed that a simple power-law behavior of the conductivity
$\sigma(E)$ as in Eq.\ (\ref{dc_cond}) was valid even for $E \ll E_c$
and $E \gg E_c$. Furthermore, we assumed that it is enough to
consider an averaged density of states $\rho(E)$. While the first
assumption is of course crucial, the second assumption is of less
importance as we have checked: Local fluctuations in $\rho(E)$ will
lead to fluctuations in the thermoelectric properties for finite $T$,
but do not lead to a different $T\rightarrow 0$ behavior: $S$ remains
finite with values as given in Tab.\ \ref{tab:nu}. Moreover, averaging
over many samples yields a suppression of these fluctuations and a
recovery of the previous behavior for finite $T$.
In this context, we remark that --- naively assuming all other parts
of the derivation are unchanged --- implications of many-particle
interactions such as a reduced single-particle density of states at
$E_F$ \cite{coulombgap}, will only modify the $T$ dependence of $\mu$.
Consequently, the $T$ dependencies of $S$, $\sigma$, $K$, and $L_0$
may be different, but their values at the MIT remain the same.
Our results also suggest that the critical regime is very small.
Namely, as the filling increases slightly from $n= 97.74\%$ to
$97.80\%$, the behavior of the system changes from metallic to
critical and finally to insulating. Up to the best of our knowledge,
such small changes in the electron concentration have not been used in
the measurements of $S$ as in Refs.\
\cite{lauinger,sherwood,lakner}. We emphasize that such a fine tuning
of $n$ is not essential for measurements of $\sigma$ as is apparent
from Fig.\ \ref{dccondfig}.
Of course, one may also speculate \cite{enderby} that these results
suggest that a true Anderson-type MIT has not yet been observed in the
experiments.
\begin{acknowledgement}
We thank Frank Milde for programming help and Thomas Vojta for
useful and stimulating discussions. We gratefully acknowledge
stimulating communications from John E.~Enderby and Yoseph Imry.
This work has been supported by the DFG as part of SFB393.
\end{acknowledgement}
|
\section{INTRODUCTION}
Theoretical modeling of chemical abundance gradients in galaxies provides
an important tool for understanding galactic evolution. However, chemical
evolutionary models applied for the dynamics of galaxies on a Hubble time
scale encounter some principal difficulties (Shore \& Ferrini 1995).
The chemical evolution of galaxies is usually considered in the one-zone
approximation with a postulated accretion rate and some assumed law of
star formation. These models do not take into account global hydrodynamic
processes of the redistribution of the matter in the galactic disks.
The spiral density waves, viscosity, and the disk-halo-disk circulation
work in different ways to redistribute matter within the galactic
disks during the galactic evolution, which considerably complicates the problem.
We discuss one class of galaxies where the modern theories of the chemical
evolution can be successfully applied. These are the starburst galaxies
with the large scale rings of an enhanced star formation. These galaxies
are believed to be the result of recent galactic collisions, and result
from the passage of a compact companion galaxy through the disk of a galaxy
along its rotation axis (Lynds \& Toomre 1976). Such collisions produce
the wave of enhanced star formation, and are responsible for the nature of
the large scale rings of the young massive stars observed in some galaxies.
The time-scale of bursts of star formation in the galaxies similar to the
Cartwheel is much shorter compared to the Hubble time, making secular
hydrodynamic processes unimportant. The chemical evolution in the fast
waves of star formation is governed therefore by a few parameters which
can be measured, or at least estimated, and ring galaxies provide a
possibility to determine the relative role of SNIa and SNII in the
enrichment of the interstellar matter by heavy elements.
It is known that there are two physically different sources of heavy
element production. The $\alpha$-elements and oxygen
are mainly produced in massive stars experiencing SNII explosions
(Woosley \& Weaver 1995), and iron peak elements
are substantially contributed by SNIa which are assumed
to be the end product of close binary evolution (Tsujimoto et al. 1995).
The relative role of these mechanisms is somewhat unclear due to
uncertainties in knowledge of the input parameters, and uncertainties of
stellar evolutionary models (Ishimaru \& Arimoto 1997; Gibson, Loewenstein \&
Mushotsky 1998).
The clock of the SNIa explosions and thus the enrichment due to SNIa is
substantially delayed with respect to SNII. Such a delay in iron chemical
enrichment compared to the $\alpha$-elements will reveal itself in the
star-forming ring galaxies as a relative abundance $[Fe/\alpha]$ radial
gradient. The value of this gradient depends only on a few parameters
such as the velocity of the wave, the slope of the initial mass function,
and on the fraction of close binary systems producing iron via type SNIa
explosions.
The number of ``free'' parameters is hence drastically reduced, and
ring galaxies give us unique opportunity to make observational conclusions
about the role of SNIa explosions in chemical evolution.
\section{CHEMICAL EVOLUTION MODEL}
As in our previous paper (Korchagin et al. 1998), we discuss two possible
mechanisms of the formation of large-scale star-forming rings.
The conventional scenario assumes that rings originate in the direct
collisions of disk galaxies with their companions.
Instead, we consider the possibility that galactic collision only
plays the role of a ``detonator'' stimulating further self-propagating
star formation.
This model assumes that the wave is a self-organized phenomenon,
and is similar to the ``fire in the forest'' models discussed earlier.
The most observationally studied ring galaxy is the Cartwheel,
and we adopt therefore parameters of the model obtained for this
archetype galaxy.
The velocity of the wave was chosen equal to 90 km/s --- the value, which
gives the best fit to the optical surface brightness gradients
observed in the Cartwheel galaxy (Korchagin et al. 1998).
In both models we use the Salpeter
IMF ($\alpha = 1.35$) with masses of stars located in the interval
0.5~M$_\odot \leq m_s \leq 50$~M$_\odot$.
It is remarkable however, that the ratio of iron to oxygen
production in the wave does not depend on the parameters determining
the rate or the efficiency of star formation. This ratio is mainly determined
by the IMF slope, and by the heavy element output of the SN
explosions. This fact makes the theoretical predictions of the radial
dependence $[Fe/O]$ ratio quite robust.
The mathematical formulation of both models, as well as
the parameters prescribing the rate of star formation
can be found in Korchagin et al. (1998).
The basic ideas of the chemical evolution model which we
incorporate here to describe abundance gradients in the
starburst ring galaxies were outlined by Tenorio-Tagle (1996).
The process of element enrichment in galaxies can be summarized
as follows. The freshly formed elements ejected by the SN
explosions fill the hot interiors of the bubbles,
which were created earlier by the stellar winds, and/or
by the release of energy in a solitary, or multiple SN explosions.
The new elements are mixed with pre-existing matter
through evaporation of a part of the cold shell surrounding the bubble,
before they are expelled into the halo of the galaxy in a hot,
SN-driven winds (``chimneys''). Hot enriched gas eventually
cools and ``rains'' back on the galactic disk completing the
circle of the element enrichment in a new event of star formation.
Recent observations support this picture. Kobulnicky (1999) did not find
any evidence for metal abundance enhancement in the vicinity of young
star clusters, concluding that the freshly ejected materials are stored
in a hot, $10^6$ gas phase.
Elmegreen (1997), considering processes of mixing and
contamination of interstellar gas after passage of a spiral density wave,
came to a similar conclusion. He finds that
there is no element mixing on time scales of a few $\times 10^7$ years.
The following processes determine spatial distribution of heavy elements
in the wake of a star-forming wave.
{\it a) Stellar Evolution:} The new-born stars ``inherit'' heavy element
abundances of the pre-existing gaseous disk, and hence their atmospheres
are not expected to contain the heavy elements produced in the current
starburst. Gas released by the low-mass single stars during their red
giant phase and the SN explosions with mass $m<11$ is assumed to have
oxygen and iron abundances equal to that at the birth epoch of the stars.
We take this abundance as to be one fifth of the solar.
The fraction of mass returned to the
interstellar medium by stars with $m<11$ during their evolution is taken
from K\"oppen \& Arimoto (1991).
Stars with masses $m > 11 M_{\odot}$ experience SNII explosions
and release freshly formed elements. We use in our simulations the
oxygen and iron yields taken from
models of Woosley \& Weaver (1995) and Nomoto et al. (1997).
The SNIa result from C-deflagration of white dwarfs in binary
systems with masses of progenitors $\le 8 M_{\odot}$.
The nature of the SNI progenitors is rather unclear, and different scenarios
give the lifetimes of SNIa progenitors in the interval $\sim 0.1 - 0.6$ Gyr
(Branch 1998).
In our simulations, we consider therefore three different mass intervals
for the SNIa progenitors.
Namely, we assume that SNIa are produced in binaries with primary
stars of 2.5--8 $M_{\odot}$, 2.5--6 $M_{\odot}$, and 2.5--3.5 $M_{\odot}$.
Hence the first supernovae explode at $4\times10^7$, $7\times10^7$ and
$2.66\times10^8$ years after the start of the burst, for the three mass
intervals. In the latter case, SNIa become unimportant in the heavy element
enrichment by fast ring waves of star formation such as inferred in Cartwheel.
We adopt the nucleosynthesis output for SNIa from the updated
W7-model of Thielemann, Nomoto \& Hashimoto (1993) with fraction of binaries
$f_{SNI}$ equal to 0.04 --- the value which Kobayashi et al. (1998) found to be
adjusted to reproduce the chemical evolution in the solar neighborhood.
{\it b) Evaporation:}
There are two sources of mixing of the chemically unprocessed gas
with the freshly formed elements inside the hot bubble.
The first mechanism is the classical thermal conductivity, which results
typically a few thousand solar masses being evaporated from the shell
surrounding the hot bubble
(e.g. Shull \& Saken 1995). A larger potential source of mixing,
but far more uncertain, is the photo-evaporation or shock
ablation of the ambient clumpy medium. We consider therefore the following
models of the chemical enrichment in the Cartwheel.
In the first model, we assume that
the freshly formed elements are instantaneously mixed with the remaining
ambient gas, or equally, that all remaining ambient gas is transformed
into a hot phase by the star formation process. In the second model,
we consider the conductive evaporation of shells to be the source of
mass in the bubble. For comparison, we also computed the $[Fe/O]$ gradients
for pure SN ejecta without mixing.
Using expression for the conductive mass-loss of the shell in the adiabatic
bubble (Shull \& Saken 1995) and assuming $10^{51}$ erg of energy released by
each SN and the shell expansion time of $10^6$ yrs, one can write the
expression
for the SN-evaporated mass in units $10^7 M_{\odot}$, $10^6$ yr and 1 kpc as:
\begin{equation}
M_{evap} = (0.7\times10^{-4})
n_0(r,t)^{-2/35} \kappa^{2/7}
\end{equation}
Here, $n_0(r,t)$ is the density of the ambient gas, and $\kappa < 1$
is a scaling factor.
{\it c) Cooling:} The hot processed gas streams out in the halo,
and the subsequent abundance gradients are determined by the
concurrent processes of cooling and diffusion. Using Raymond's et al. (1976)
cooling coefficient we find that the cooling time of hot X-ray emitting
gas with the temperature $\sim$ few $\times 10^6 K$ and the density
$10^{-2}-10^{-3}$~cm$^{-3}$ is about a few $\times (10^7 - 10^8)$ yrs which
is large, or comparable to the dynamical time of the starburst.
At the temperature interval typical for the X-ray emitting halo gas,
the heavy elements, particularly iron, dominate
the cooling process, accounting from half to three-forth of the total
cooling rate (Raymond et al 1976). The heavy element abundances in the
Cartwheel are deficient by factor of ten as compared to the Orion Nebula
in our galaxy (Fosbury \& Hawarden 1977), which will increase the
estimated cooling time. We assumed therefore
that cooling is unimportant during the starburst time in the Cartwheel.
{\it d) Diffusion:} The heavy ion admixtures diffuse into a hot hydrogen
plasma of halo, decreasing possible abundance gradients produced by the
SN explosions. The diffusion of light ions is effective on the
time scale of the starburst (Tenorio-Tagle 1996), but is ineffective
for heavy, highly ionized ions. In plasma with temperature of a
few $\times 10^6$ K, the oxygen will be ionized up to OVI, and iron up to
FeXI--FeIV. The mean free paths of oxygen and iron in plasma with
temperature $5\times 10^6$ K and density $10^{-3}$ will be about 50pc
and 12pc correspondingly, which gives values of the dimensionless diffusion
coefficients $D_O \approx 5\times 10^{-3}$ and $D_{Fe} \approx 6\times 10^{-4}$.
Our simulations show, that with such values, the diffusion of heavy
elements is unimportant on the time-scales of passage wave front in the disk.
{\it e) Grain Formation:} For the pressure typical in a SN-processed
gas, the grain formation is effective for temperatures below 1000 K
(Gail \& Sedlmayr 1985). Hot halo gas will preserve therefore the abundance
compositions which were formed during the ejecta.
However, if the hot gas had enough time to cool, iron will be selectively
depleted with respect to oxygen due to grain formation.
Assuming that the SN ejecta mix with all the remaining ambient gas,
we can write the equation describing the spatial distribution of the heavy
elements in the wake of the star-forming wave:
\begin{eqnarray*}
{\partial M_i(r,t) \over \partial t } =& -Z_i^{0} B(r,t)
+ D_i {\partial^2 M_i \over \partial r^2} \\
+&
\int_{M_{min}}^{M_{max}} B(r,t-\tau_{m}) \phi (m) R_i(m)dm
\quad (2)
\end{eqnarray*}
Here $B(r,t)$ denotes the birth rate of stars as a function of time and
radius of the galactic disk, $M_i$ is the surface density of an i-th element,
$Z_i^{0}=M_i^0/M_{GAS}$ is the initial abundance of the i-th element in the
star-forming gas $\phi (m)$ is
the initial mass function with minimum of maximum masses $M_{min}$ and
$M_{max}$ respectively. $D_i$ is the diffusion coefficient
of i-th element, and $M_{GAS} $ is the initial constant surface
density of gas. The coefficient $R_i(m)$ gives ratio of mass of the
ejected element to the mass of a star $m$.
In a similar way, we can write the equation governing the chemical evolution
behind the wave
when the stellar ejecta are mixed with gas thermally evaporated from
the shells with its mass prescribed by the equation (1).
Once the IMF is fixed, the birth and death rates can be determined at each
radial grid zone
making use of the lifetimes of stars. Then the
heavy element distributions can be computed by means of equations (1) and (2).
\section{Results }
Figure 1 illustrates the $[Fe/O]$ enrichment produced by the wave of
star formation for the three different scenarios of the element mixing and
three different mass intervals of the pre-SNIa stars. The position of the wave
taken at time $t=240 Myr$ corresponding to the present location of the
outer ring
in the Cartwheel galaxy. The left frames show the abundance gradients for the
Nomoto et al. (1997) SNII outputs, and the right frames present the results
of simulations for the Woosley \& Weaver (1995) SNII iron and oxygen outputs.
The bottom frames of Figure 1 show the $[Fe/O]$ profiles if masses
of pre-SNIa stars 2.5--3.5 $M_{\odot}$. In this case, all SNIa progenitors have
lifetimes longer than the time of the wave propagation , and the role of SNIa
stars in the enrichment is negligible. The solid lines of the Figure 1
show the $[Fe/O]$ radial dependence when hot SN ejecta are mixed with the
part of ambient gas evaporated from the SN shells. For comparison,
we also plot the abundance gradients when the freshly synthesized elements are
instantaneously mixed with the all ambient gas (dashed lines), and the
abundance gradients produced by the pure SN ejecta (dotted lines).
Figure 1 depicts more sharp growth of the oxygen abundance as compared
to the iron one in front of the wave of star formation.
Such behavior directly follows from the
yields of iron and oxygen by the SNII explosions. In the model of Woosley \&
Weaver (1995) the oxygen yield increases with stellar mass until the progenitor
mass reaches approximately 40 $M_{\odot}$, and then tends to saturate.
Supernova models of Nomoto et al. (1997) give rapid increase of oxygen yield
with the mass of progenitor. The iron yield on the contrary decreases with
stellar mass in both the models. Therefore, SNII with moderate masses of
progenitor are responsible mainly for the iron production, whereas SNII
with progenitors of higher mass produce mainly oxygen.
Large $[Fe/O]$ gradients in the region of the wave of star formation are
changed to slower growth of the $[Fe/O]$ ratio towards the center of galaxy
when metal enrichment due to SNIa events start to become important.
The top and intermediate frames of Figure 1 show the radial profiles
of $[Fe/O]$ for the SNIa progenitors with masses 2.5--8 $M_{\odot}$ and
2.5--6 $M_{\odot}$, respectively. The ``pollution'' by the SNIa hot ejecta
causes the increase of the $[Fe/O]$ gradient towards the center of the disk.
The effect is obviously most noticeable for the pure SN ejecta (dotted lines).
The element mixing diminishes the effect, which however remains
about 0.05 dex in the Fe/O ratio in both SNII output models and both
models of mixing of newly-formed elements.
The radial abundance profiles of oxygen and iron obtained for the density
wave are much similar to those obtained for the induced wave of star formation.
The radial distributions of relative abundance $[Fe/O ]$ are qualitatively
independent of the particular model of star formation and
the uncertain parameters such as the rate of
star formation and the ``efficiency'' of star formation.
This makes our conclusions robust, and the relative
abundance gradients $[Fe/\alpha ]$ might be used as a tool for
determining the relative role of SNIa in the chemical enrichment.
\section{DISCUSSION}
The nebular abundances measured by
Fosbury \& Hawarden (1977) in the Cartwheel
pertain to the heavy element abundances before the passage of the wave.
Therefore, observations
in the inner ring regions are required to test our predictions.
Interstellar abundances can be inferred however using X-ray emission
from hot halo gas.
The launch of the Advanced X-ray Astrophysics Facility ($\it AXAF$)
with its spatial resolution of $\approx 0.^{\prime \prime}5$ would
allow to measure the heavy element gradients in the ring galaxies
of star formation.
If the hot gas has enough time to cool after the passage of the wave
of star formation, and the newly-formed elements reside in the warm
ionized or the cold phases of interstellar medium, then abundances
can be inferred using metallic absorption lines
towards background objects. The absorption lines of most atoms found in the
interstellar medium occur at ultraviolet wavelengths, and hence
abundance determinations require satellite observations.
With the launch of {\it Hubble Space Telescope} interstellar abundances
are now available for a number of atoms, including iron and the $\alpha$
elements $O, N, C, Mg$ etc. (Savage \& Sembach 1996).
In external galaxies, the measurement of interstellar abundances depends
on the availability of background sources. The most widely
used background sources are quasars, using which abundances in halos of
galaxies and in the intervening clouds have been determined (e.g. Morton
et al. 1980). In a recent study, Gonz\'alez-Delgado et al. (1998) have
detected interstellar absorption lines towards the bright starburst
nucleus of NGC\,7714 in the ultraviolet spectrum taken with the {\it
Hubble Space Telescope}. Hence, when background sources are available the
present technology allows for the measurement of interstellar absorption lines.
In ring galaxies, the measurement of interstellar abundance thus depends
on the inclination of the ring to the line of sight. In cases, where the
ring is favorably aligned, the spectrum of center regions and the ring HII
regions are likely to contain interstellar absorption lines.
\begin{acknowledgments}
VK acknowledges Prof. S. Miyama for hospitality, and National Astronomical
Observatory of Tokyo for providing COE fellowship.
\end{acknowledgments}
\newpage
|
\section{Introduction}
The Short Wavelength Spectrometer (SWS) on board the Infrared Space
Observatory (ISO) has allowed us, for the first time, to observe the
pure rotational spectrum of H$_2$ and study the OTP
ratio in warm (T$\sim$ 100 - 1000 K) regions.
Previous to ISO observations, the OTP ratio of H$_2$ had been
estimated
towards regions heated by UV radiation and shocks
to temperatures $>$ 2000 K
using the relative strengths of the
vibrational lines. In outflow regions where shock excitation
is the main heating mechanism, observations of the vibrational
lines typically reveal OTP ratios of $\sim$ 3 (Smith,
Davis \& Lioure 1997). This is the expected value in thermodynamic
equilibrium for gas with temperatures $>$ 200 K.
However, towards the regions heated mainly by the UV radiation
(PDRs), the vibrational lines give OTP values in the range 1.5 - 2
(see e.g. Chrysostomou et al. 1993).
The pure rotational lines of H$_2$ have been detected by ISO toward
several Galactic regions (S140: Timmermann et al. 1996; Cepheus A West:
Wright et al. 1996; BD+40$\deg$4124: Wesselius et al. 1996;
HH54: Neufeld et al. 1998).
The observations towards all these regions except HH54
are consistent with the
H$_2$ rotational lines arising in gas with temperature $>$ 200 K
and with an OTP ratio of 3. This value is not consistent with those
derived from the vibrational lines in dense PDRs
but is in agreement
with theoretical predictions.
Recently Sternberg \& Neufeld (1999) have argued that the low
OTP ratios measured from vibrational lines
do not represent the actual OTP ratio in PDRs
but it is simply a consequence of the optical depth effects in the fluorescent
pumping of the vibrational lines.
The non-equilibrium OTP ratio measured towards HH54
has been interpreted as arising in hot-shocked excited gas that has not
reached the equilibrium yet. In this interpretation, the observed
OTP ratio would be the legacy of an earlier stage in the thermal history of the
gas when the gas temperature was 90~K.
In this {\it Letter} we report the observations of the H$_2$ rotational lines
towards the prototypical PDR associated with the reflection nebula
NGC 7023. An OTP ratio in the range 1.5 - 2 is derived from
our observations. This is the first
detection of an OTP ratio $<$ 3 in a PDR based on the
pure H$_2$ rotational lines.
\section{Observations}
The observations of the S(0), S(1), S(2), S(3), S(4) and S(5) rotational
lines of H$_2$ towards the peak of the PDR associated with
NGC 7023 (R.A. (2000):\ra[21 01 32.5], Dec(2000):\dec[68 10 27.5]).
were made using the Short-Wavelength Spectrometer (SWS)
(de Grauuw et al. 1996) on board the Infrared Space Observatory
(ISO) (Kessler et al. 1996) during revolution 514 with a total on-target time
of 5149 s. At the spectral resolution of this mode
($\lambda$/$\Delta$$\lambda$ $\sim$ 1000-2000) all the observed lines are
unresolved. Data reduction are carried out with version 7.0 of the Off Line
Processing routines and the SWS Interactive Analysis at the ISO Spectrometer
Data Center at MPE. Further analysis has been made using the ISAP software
package. The uncertainities in the calibration are of 15 \%,
25 \%, 25 \%, 25 \%, 20\% and 30\% for the S(5), S(4), S(3), S(2), S(1) and
S(0) lines respectively (Salama et al. 1997).
\section{H$_2$ rotational lines}
In Fig. 1 we present the spectra of
the S(0), S(1), S(2), S(3), S(4) and S(5) H$_2$ lines towards the
peak of the PDR associated with NGC 7023. The observed intensitites
and some interesting observational parameters are presented in Table 1.
The data have been corrected for dust attenuation
using the extinction curve of Draine \& Lee (1984),
and a value of 0.43
for the dust opacity at 0.55 $\mic$.
This value has
been derived from the ISO LWS01 spectrum (Fuente et al.
1999, hereafter paper II).
The extinction for the
S(0), S(1), S(2), S(3), S(4) and S(5) lines is very small and
amounts to 0.01, 0.02, 0.02,
0.04, 0.02 and 0.01 respectively.
Fig. 2 shows the rotational diagram for the H$_2$ corrected for extinction
effects and assuming extended emission. (Note that the errors in Fig. 2
are entirely dominated by the calibration uncertainities.)
This diagram shows that the ortho-H$_2$ levels
have systematically lower N$_u$/g$_u$ values
(where N$_u$ and g$_u$ are the column densities and
degeneracies of the upper levels of the transitions)
than the adjacent J--1 and J+1 para-H$_2$ levels
producing a $``$zig-zag" distribution.
In fact, the ortho-H$_2$ levels seem to define
a curve which is offset from that of the para-H$_2$ levels (see Fig. 2).
This is the expected trend if the OTP ratio is
lower than 3. The offset between the two set of
data is systematic and seems to show the trend of
being larger for the low energy levels (S(0) and S(1) transitions),
than for the high energy levels (S(2), S(3), S(4) and S(5) transitions).
The offset between the ortho- and para- H$_2$ curves is larger
than the observational errors and
cannot be due to the different apertures for the different lines.
For the case of a point-like source we would have to correct by a factor
of 1.93 the intensity of the S(0) line and by a
factor of 1.35 the intensities of the S(1) and S(2) lines.
In this case, the offset between the para- and ortho- curves
would increase.
Based on the spatial distribution of the HI(21 cm) line (Fuente et al. 1998)
and those of the CII (158 $\mic$) and OI (63 $\mic$) lines
(Chokshi et al. 1988,
paper II), we will assume in the following a beam filling factor of 1
for all the observed H$_2$ lines.
Although the exact shape of the rotational diagram depends
on the excitation conditions of the region,
the $``$zig-zag" features cannot be
explained by any model which assumes an equilibrium OTP ratio.
In thermodynamic equilibrium,
the rotational temperature between an ortho- (para-) H$_2$ level
and the next J+1 level should always increase with
the energy of the upper level giving rise to a smooth curve in the
rotational diagram regardless of the temperature profile of
the region. A $``$zig-zag" distribution implies that for
some pair of levels
the rotational temperature decreases with the energy of the upper level.
One can only get this effect with a non-equilibrium OTP ratio.
We have compared our data with the models by Burton et al. (1992)
in order to estimate the incident UV field, density and OTP ratio.
Burton et al. (1992) assumed a constant value of 3 for the OTP
ratio in their calculations. The best
fit to our data is for an incident UV field of G$_o$ = 10$^4$
in units of the Habing field and a density of n = 10$^6$ cm$^{-3}$,
but it systematically underestimates
the intensity of the para-H$_2$ transitions and overestimates the
intensities of the ortho-H$_2$ transitions (see open triangles in Fig. 3).
This is the expected behavior if one assumes
an OTP ratio of 3 and the actual OTP ratio in the region is $<$ 3.
We have corrected the line intensities predicted by the model for
different values of the OTP ratio assuming that the total amount of H$_2$
molecules at a given temperature and the line ratios between levels
of the same symmetry are not affected by the OTP ratio. These
assumptions are correct for the low-J transitions where
the collisional excitation dominates.
In Fig. 3 we compare our data with the predicted diagram
for an OTP ratio of 3 which correspond to the model
without any correction (open triangles),
2 (open circles) and 1 (open squares).
The diagram for an OTP ratio of 3 clearly does not fit any of
our observational points. We find the best fit to our data with
an OTP ratio of 2.
In this case, the predicted intensities for the S(1), S(2) and S(3) lines
are in agreement with the observed values.
Only the predicted intensity for the S(0) line is not consistent
with the observations. To fit the intensity of the S(0) line it is
necessary to assume an OTP ratio of $\leq$ 1.5 (see the case OTP = 1 in
Fig. 3), but in this case, we will have
a worse fit for the S(1), S(2) and S(3) lines than for an OTP ratio of 2.
As discussed in Section 4, this suggests a possible variation of
the OTP ratio with the temperature.
The cooler gas emitting only in the S(0) line
could have an OTP ratio of $<$ 1.5, while
the warmer gas emitting in the S(1), S(2) and S(3) lines
could have an OTP ratio of 2.
However, to establish this variation
unambigously, it is necessary to have a
very accurate knowledge of the excitation
conditions in the region. To be conservative,
we will conclude that the OTP ratio
is in the range 1.5 - 2 in this PDR. Martini et al. (1997) found an
OTP ratio of 2.5$\pm$0.4
using the near-IR H$_2$ vibrational lines towards their
position 1 which is 20$''$ offset from ours.
Taking into account that because of the optical depth effects in the
vibrational lines this value is just a lower limit to the
actual OTP ratio, it proves that the OTP ratio is close to 3
for the gas with kinetic temperatures T$_k$ $>$ 2000 K.
This difference between the OTP ratios derived from the
rotational and vibrational lines
argues in favor of a variaton of
the OTP ratio with the kinetic temperature.
Since the OTP ratio in this source is different from the equilibrium
value, the rotation temperature between an ortho- and a para- level
does not represent an estimate of the gas kinetic temperature. To
estimate the gas kinetic temperature we have calculated the rotation
temperature between levels of the same symmetry.
For the para-H$_2$ levels, we have derived a rotation temperature
of $\sim$ 290 K from the S(0) and S(2) lines and of
$\sim$ 500 K from the S(2) and S(4) lines. For the ortho-H$_2$ levels,
the derived temperature is $\sim$ 440 from S(1) and S(3) lines and
$\sim$ 700 K from the S(3) and S(5) lines.
Based on these calculations we
conclude that the OTP ratio is $\sim$ 1.5 - 2 in the
gas with kinetic temperatures $\sim$ 300 - 700 K.
We derive a total H$_2$ column density of 5 10$^{20}$~cm$^{-2}$
assuming a rotation temperature of 290 K between the J = 0 and J = 2
para-levels and of 440 K between the J = 1 and J = 3 ortho-levels.
\section{Discussion}
Based on the H$_2$ rotational lines data
we have derived an OTP ratio in the range
of 1.5 - 2 for the gas with kinetic temperatures $\sim$ 300 - 700 K
towards the reflection nebula NGC 7023. This is the second object with
a non-equilibrium OTP ratio measured from the H$_2$ pure rotational lines.
The first non-equilibrium OTP ratio was
detected towards the outflow source HH54, and interpreted as arising in
shock-heated gas that has not reached
the equilibrium yet. This interpretation is not plausible for NGC 7023.
The high dust temperature,T$_d$$\approx$ 40 K, and the detection of
the SiII (34.8 $\mic$) and CII (157.7 $\mic$) lines (paper II)
proves the existence of a PDR in this region.
Although a shock component might also exist
(see Martini et al. 1997), the heating is dominated by UV photons.
It is then expected that most of the warm H$_2$ arises in the PDR.
A non-equilibrium value of the OTP ratio is not expected for the
physical conditions prevailing
in a dense PDR. The initial OTP ratio after
the H$_2$ formation is very uncertain. Because of the large exothermicity
of the formation process, it is expected to be 3 .
However, this OTP ratio can change if the H$_2$ molecules remain on
the grain surface long enough to reach the equilibrium at the dust
temperature. In this case, the OTP ratio after H$_2$ formation
will be that of the equilibrium at the grain temperature.
After the ejection of the H$_2$ molecules to the gas phase,
exchange reactions with H and H$^+$ change the OTP ratio
until achieving the equilibrium at the gas temperature.
For the gas temperatures traced by the H$_2$ rotational
lines ($>$ 200 K), the equilibrium OTP ratio
is 3. One possibility to explain
the low OTP ratio observed in NGC 7023 is to suppose that
the OTP ratio in the H$_2$ formation is lower than 3,
and once in the gas phase,
the H$_2$ molecules are destroyed before attaining the
equilibrium value at the gas temperature.
The dust temperature in a PDR with G$_o$ = 10$^4$
and n = 10$^6$ cm$^{-3}$ is $<$ 100 K. In particular, the dust temperature
measured towards the PDR peak in NGC 7023 using our LWS01 spectrum is
40 K which corresponds to an equilibrium OTP ratio of $\sim$ 0.1.
The OTP conversion in the atomic region is dominated by H - H$_2$ collisions
with a rate
of 10$^{-13}n$ s$^{-1}$ where $n$ is the atomic hydrogen density
(see Sternberg \& Neufeld 1999 and references therein).
The hydrogen density derived from the HI image published
by Fuente et al. (1998) is $\sim$ 5 10$^3$ cm$^{-3}$. A similar value
($\sim$ 0.5 - 1 10$^4$ cm$^{-3}$)
was derived by Chokshi et al. (1988) based on the OI and CII lines.
However, densities larger than 10$^5$ cm$^{-3}$ are derived from molecular
data (Fuente et al. 1996, Lemaire et al. 1996, Gerin et al. 1998).
The density of atomic hydrogen in the region with T$_k$ $\sim$ 300 K
is expected to be about an order of magnitude lower than
the H$_2$ density. Then we assume an atomic hydrogen density of
$\sim$ 10$^4$ cm$^{-3}$ in our calculations.
With this value, the
OTP conversion rate due to H$_2$ - H collisions
is 10$^{-9}$ s$^{-1}$.
At the cloud surface the unshielded photodissociation rate is
$\sim$ 5 10$^{-11} G_\circ$ s$^{-1}$ $\sim$ 5 10$^{-7}$ s$^{-1}$, i.e.,
2 orders of magnitude larger than the OTP conversion rate.
Then, in the cloud surface, before self-shielding is important
(A$_v$ $\leq$ 0.3 mag, T$_k$$>$ 500 K), an OTP ratio lower than 3
can be explained by assuming that the OTP ratio in the H$_2$ formation
is the equilibrium value at the grain temperature.
Deeper into the molecular cloud, when H$_2$ is self-shielded,
one can consider that the
H$_2$ destruction rate is similar to the H$_2$ reformation rate
which is given by $\sim$ 3 10$^{-17} n$ s$^{-1}$. In this region,
a lower limit to the OTP conversion rate is given
by the conversion rate due to H$_2$ - H$^+$ collisions
which is $\sim$ 10$^{-17} n$ s$^{-1}$ (Sternberg \& Neufeld 1999).
The OTP conversion
rate is of the same order than the destruction rate and the
OTP ratio is expected to be close to the equilibrium.
According to these estimates one expects to have an OTP ratio close to 3
at low temperatures and a non-equilibrium OTP ratio at higher temperatures.
The contrary trend is derived from our observations.
The high energy S(1), S(2) and S(3) lines are fitted with an
an OTP ratio of 2 while the low energy S(0) line is better fitted
with an OTP ratio of 1.5.
Another possibility to explain the non-equilibrium OTP value
in NGC 7023 is to consider the case of a dynamic PDR, i.e. the
dissociation front is advancing into the molecular cloud. In this case
we do not have to assume a non-equilibrium OTP ratio in
the H$_2$ formation.
The PDR is being fed continuously by the cool
gas of the molecular cloud
in which the equilibrium value of the OTP ratio is lower than 3.
This gas is heated by
the stellar UV radiation to temperatures $>$ 200 K but leaves the PDR
before attaining the equilibrium OTP ratio at this temperature.
In this scenario,
the gas which is expected to have an OTP
ratio smaller than the equilibrium value is the gas that has
more recently been
incorporated into the PDR, which is also the gas at
lowest temperature.
The OTP ratio is expected to increase and reach values
close to 3 at high temperatures.
This behavior is consistent with the
trend observed in our data.
As discussed above,
the OTP H$_2$ conversion
is mainly due to H - H$_2$ collisions with a conversion rate
of $\sim$ 10$^{-9}$ s$^{-1}$.
To have a significant fraction of the gas with a non-equilibrium
OTP ratio the photodissociation front must advance about
1 - 2 mag in the conversion time, i.e.,
it must penetrate into the molecular cloud
at a velocity of $\sim$ 10$^7 n^{-1}$ kms$^{-1}$.
Skinner et al. (1993) proposed the existence
of an anisotropic ionized wind associated with this star based on radio
continuum observations. Later, Fuente et al. (1998) detected
an HI outflow with a velocity of
$\sim$ 7.5 kms$^{-1}$. They proposed that this outflow is formed
when the gas that has been photodissociated by the UV radiation,
is accelerated by the stellar winds along the walls of a biconical
cavity. The velocity of the
HI outflow cannot account for the velocity
at which the photodissociation front must advance into the
molecular cloud to have an OTP ratio lower than 3
unless the OTP conversion rates are severely
overestimated (by a factor of $>$ 10) and/or the density of the
outflow is $>$ 10$^5$ cm$^{-3}$. In spite of
this problem, a photodissociation front penetrating into the
cloud because of the outflow seems to be the
most plausible explanation for the
low OTP ratio measured in this region.
The existence of an outflow can also explain
the difference between this region and other PDRs like S140 in which
an OTP ratio of 3 has been
derived from the H$_2$ rotational lines.
\acknowledgments
We would like to thank the referee for his/her helpful comments.
This work has been partially supported by the Spanish
DGES under grant number PB93-0048 and the Spanish PNIE under
grant number ESP97-1490-E.
N.J.R-F acknowledges Conserjer{\i}a de Educaci\'on y Cultura de la
Comunidad de Madrid for a pre-doctoral fellowship.
\clearpage
\newpage
|
\section{Introduction}
With new instruments such as MEGACAM at CFHT (\cite{Bouetal98}\ 1998) and the
VST at the European Southern Observatory (\cite{Arnetal98}\ 1998),
wide-field surveys detect the weak lensing of faint galaxies by large scale
structure will soon become a reality (see \cite{Mel98}\ 1998 for a
recent review). Weak lensing by large-scale structure produces
a correlated distortion in the ellipticities of the galaxies on
the percent level (\cite{Blaetal91} 1991; \cite{Mir91}\ 1991;
\cite{Kai92} 1992) which can be used to measure
a two-dimensional projection of the intervening mass distribution
(\cite{TysValWen90} 1990; \cite{KaiSqu93} 1993).
If the redshift of the source galaxies are known, then more information
can be extracted out of weak lensing by tomography, i.e. differencing
the two-dimensional projected images to recover the three-dimensional
distribution. In the absence of spectroscopy, approximate redshifts for
the faint galaxies can be determined through photometric techniques
(see e.g. \cite{Hogetal98} 1998 and references therein) and with the large
number of galaxies at $R \lesssim 25$
($\sim 10^5$ deg$^{-2}$) the properties of
distribution can be known to good accuracy (\cite{Sel98} 1998).
Indeed the weak lensing surveys
already plan to use photometric redshift information at least on a small
subsample to measure the low order
moments of the distribution such as its mean. These are important for
determining the cosmological implications of the data (\cite{Smaetal95a} 1995;
\cite{ForMelDan96} 1995; \cite{LupKai97} 1997).
The potential of tomographic techniques, especially in the wide-field
limit where the cosmological information is completely contained in
the two-point functions or power spectra, remains largely unexplored.
Indeed most studies of weak lensing (e.g.
\cite{JaiSel97} 1997;
\cite{Kai92} 1998;
\cite{HuTeg99} 1999)
simply assume a delta-function distribution of galaxies
making tomography impossible.
In this {\it Letter}, we study the prospects for weak lensing tomography
within the framework of the adiabatic cold dark matter (CDM) class of models for
structure formation.
We begin by defining the power spectrum statistics
for an arbitrary set of galaxy redshift distributions.
These are the power spectrum of the convergence map for each distribution
and the cross-correlation between the maps. We then quantify
how much additional information can be extracted by subdividing a single
magnitude-limited sample into bins in redshift and analyzing
their joint power spectra and cross-correlation. We conclude with
a discussion of how errors in photometric redshifts might affect
tomographic techniques.
\section{Power Spectra}
Generalizing the results of \cite{Kai98} (1992, 1998), we
define the angular power spectra and cross-correlation of sky maps
of the convergence based on a series of galaxy redshift distributions
$n_i(z)$ (see also \cite{Sel98} 1998)
\begin{eqnarray}
{{\bf P}^{\kappa}_{ij}}(\ell) &\equiv& {1 \over 2\ell+1}
\left< a^*_{(\ell m)\, i} a_{(\ell m) \, j} \right> \, \nonumber\\
&\approx& 2\pi^2 \ell \int d D \, {{g_i(D) g_j(D)} \over
D_A^3(D)}
\Delta_\Phi^2 (k_\ell,D) \,,
\label{eqn:ppsi}
\end{eqnarray}
where $a_{(\ell m)\, i}$ are the spherical harmonic coefficients of
the maps. Here
$D = \int_0^z (H_0/H) dz$ is the dimensionless comoving
distance and
\begin{equation}
D_A(D)=
\Omega_K^{-1/2} {\sinh(\Omega_K^{1/2}D)}
\end{equation}
is the dimensionless angular diameter distance, where
$\Omega_K=1 - \sum_i \Omega_i$ is the effective density in spatial curvature in
units of the critical density.
The efficiency
with which
gravitational potential fluctuations $\Phi$, as measured by their dimensionless
power per logarithmic interval $\Delta_\Phi^2 \equiv k^3 P_\Phi/2\pi^2$, lens
the given galaxy distribution $n_i$ is described by
\begin{equation}
g_i(D) = D_A(D)
\int_D^\infty dD'\, \left[ n_i { {dz \over dD} }\right](D') {D_A(D'-D) \over
D_A(D')} \,.
\end{equation}
Note that $n_i$ is normalized
so that $\int_0^\infty dz \, n_i(z)=1$.
Finally
$k_\ell = \ell H_0 / D_A(D)$
is the wavenumber which
projects onto the angular scale $\ell$ at distance $D$.
For small fields of view, the spherical harmonics of order $\ell$ can be
replaced by Fourier modes with angular frequency $\omega$.
We use the \cite{PeaDod96} (1996) scaling relations to evalute
$\Delta_\Phi^2$ in the non-linear density regime. Equation~(\ref{eqn:ppsi})
assumes that the redshift distributions are sufficiently wide to encompass
many wavelengths of the relevant fluctuations ($2\pi/k_\ell$)
along the line of
sight so that the Limber equation holds even tomographically
(see \cite{Kai98} 1998).
These power spectra define the cosmic signal. Shot noise in the
measurement from the intrinsic ellipticity of the galaxies adds white noise
to
the cosmic signal making the observed power spectra
\begin{equation}
{\bf C}_{ij}(\ell) = {{\bf P}^{\kappa}_{ij}}(\ell) +
{\left< \gamma_{\rm int}^2\right>\delta_{ij}/\bar n_i} \,,
\label{eqn:covariance}
\end{equation}
where $\left< \gamma_{\rm int}^2 \right>^{1/2}$ is the rms intrinsic
shear in each component, and ${\bar n}_i$ is the number density of the galaxies
per steradian on the sky in the whole distribution $n_i(z)$.
\begin{figure}[t]
\centerline{\epsfxsize=3.5truein\epsffile{fig1.eps}}
\caption{Subdividing the source population. Partitioning the galaxies
by the median redshift (or distance $D$) yields
lensing efficiencies with strong overlap.}
\label{fig:g}
\end{figure}
The distributions $n_i(z)$ need not be physically distinct galaxy populations.
Consider a total distribution $n(z)$ with
\begin{equation}
\left[ n {d z \over dD} \right](D)
\propto D^\alpha \exp[-(D/D_*)^\beta] \,,
\label{eqn:distribution}
\end{equation}
which roughly approximates that of a magnitude-limited survey,
and take $\alpha=1, \beta=4$ for definiteness (assumed throughout
unless otherwise stated). One can {\it subdivide}
the sample into redshift bins to define the distributions $n_i(z)$.
The power spectra for cruder partitions
can always be constructed out of finer ones: if the
$j$ and $k$ bins are combined, then
\begin{eqnarray}
\bar n_{j+k}^2 {\bf P}_{(j+k)(j+k)}^\kappa &=& \bar n_{j}^2
{\bf P}_{jj}^\kappa + 2 \bar n_j \bar n_k {\bf P}_{jk}^\kappa
+ \bar n_k^2 {\bf P}_{kk}^\kappa \,, \nonumber\\
\bar n_{j+k} {\bf P}_{i(j+k)}^\kappa
&=& \bar n_j {\bf P}_{ij}^\kappa + \bar n_k {\bf P}_{ik}^\kappa \,.
\end{eqnarray}
In Fig.~\ref{fig:g}, we show an example where the galaxies
with $z< z_{\rm median}$ are binned into $n_1$ and the rest into $n_2$.
Here and throughout we will take our fiducial
cosmology as an adiabatic CDM model with matter density $\Omega_m=0.35$,
dimensionless Hubble constant $h=0.65$, baryon density $\Omega_b=0.05$,
cosmological constant $\Omega_\Lambda=0.65$, neutrino mass $m_\nu=0.7$ eV,
the initial potential power spectrum amplitude $A$, and tilt $n_S=1$.
We also plot in Fig. \ref{fig:g} the lensing efficiency
function $g_i(D)$. Notice that despite having non-overlapping source
distributions (upper panel), the lensing efficiencies strongly
overlap (bottom panel) implying that the resulting convergence maps will have a
correspondingly large cross correlation. This is of course because
the high and low redshift galaxies alike are lensed by low-redshift
structures. Also for this reason, there will be always be a stronger
signal in the high redshift bins. This fact will be important for
signal-to-noise considerations in choosing the bins.
All of these properties can be seen in Fig.~\ref{fig:power}, where we plot the
resultant power spectra and their cross correlation for
the equal binning of Fig.~\ref{fig:g}.
\begin{figure}[t]
\centerline{\epsfxsize=3.5truein\epsffile{fig2.eps}}
\caption{Power spectra and cross correlation for a subdivision in
two across the median redshift $z_{\rm median}=1$
and errors for a survey of $5^{\circ}$ on
the side, $\left< \gamma_{\rm int}^2 \right>^{1/2}=0.4$, and
$\bar n = 2 \times 10^5$ deg$^{-2}$.
Note the strong correlation ${\bf R}_{ij}$ between the two power spectra make
the combination of the power spectra less constraining than a naive
interpretation of the individual errors would imply.}
\label{fig:power}
\end{figure}
\section{Redshift Binning and Parameter Estimation}
While subdividing the sample into finer bins always increases the amount of
information, there are two considerations that limit the effectiveness of redshift
divisions. The first is set by the shot noise from the intrinsic ellipticities
of the galaxies. Once the number density $\bar n_i$ per bin
is so small
that shot noise surpasses the signal in equation (\ref{eqn:covariance}), further
subdivision no longer helps. The point at which this occurs depends on the
angular scale of interest. The greater number of galaxies encompassed by the
larger angular scales boosts the signal to noise (see Fig.~\ref{fig:power} and
\cite{Kai92} 1992). Based on this criterion, one should separately subdivide the
data to extract the maximal large and small angle information.
However there is a second consideration.
If the lensing signal does not change significantly
across the redshift range of the whole distribution, then subdivision
will not add information. These considerations can be quantified by
considering the correlation coefficient between the power spectra of the subdivisions:
${\bf R}_{ij} = {\bf P}^\kappa_{ij}/
({\bf P}^\kappa_{ii} {\bf P}^\kappa_{jj})^{1/2}$.
For the model of Fig.~\ref{fig:power},
the power spectra are highly correlated (${\bf R}_{12} \sim 0.8$) even with only
two subdivisions. Thus even though
there is enough signal to noise to subdivide the sample further, one gains little
information by doing so.
One can combine these two considerations by diagonalizing the covariance
matrix and considering the signal to noise in the diagonal basis. The appropriate
strategy for subdivision depends on the true redshift distribution of the
galaxies and the model for structure formation. One should therefore
perform this test on the actual data to decide how to subdivide the sample.
Nevertheless, to make these considerations more concrete, let us consider the
specific goal of measuring the
cosmological parameters $p_\alpha$ assuming that the underlying
adiabatic
CDM cosmology described above is correct.
The Fisher information matrix can be used to quantify
the effect of subdivision. It is defined as
\begin{equation}
{\bf F}_{\alpha\beta} = -\left< \partial^2 \ln L \over \partial p_\alpha \partial p_\beta
\right>_ {\bf x} \,,
\end{equation}
where $L$ is the likelihood of observing a data set ${\bf x}$
given the true parameters
$p_1 \ldots p_\alpha$.
Generalizing the results of \cite{HuTeg99} (1998) to multiple
correlated power spectra, we obtain\footnote{When
taking these derivatives the redshift distribution $n_i(z)$ is held fixed
as opposed to the distance distribution $[n_i(z)dz/dD]$ in
equation~(\ref{eqn:distribution}).}
\begin{equation}
{\bf F}_{\alpha\beta} = \sum_{\ell=2}^{\ell_{\rm max}}
({\ell+1/2}) f_{\rm sky} {\rm tr} [{\bf C}^{-1} {\bf C}_{,\alpha}
{\bf C}^{-1} {\bf C}_{,\beta}]\,,
\label{eqn:Fisher}
\end{equation}
under the assumption of Gaussian signal and noise,
where $f_{\rm sky}$ is fraction of sky covered by the survey, the covariance matrix
${\bf C}$ was defined in equation~(\ref{eqn:covariance}), and commas denote partial
derivatives with respect to the cosmological parameters $p_\alpha$.
We take $\ell_{\rm max}=3000$ to approximate the increased covariance
due to the nonlinearities producing non-Gaussianity in the signal
(\cite{ScoZalHui99} 1999).
Since the variance of an unbiased estimator of a
parameter $p_\alpha$ cannot be less than $\sigma(p_\alpha)=({\bf F}^{-1})_{\alpha\alpha}$,
the Fisher matrix quantifies the best statistical errors on parameters possible with
a given data set.
For the purposes of this work, the
absolute errors on parameters are less relevant than the
improvement in errors from subdividing the data (see
\cite{HuTeg99} for the former).
We therefore test a 4 dimensional subset of the adiabatic CDM parameter space
to see how subdivision
can help separate initial power ($A$)
from the various contributors to the redshift-dependent evolution of power
($\Omega_\Lambda$,$\Omega_K$,$m_\nu$).
For reference the standard errors $\sigma_\alpha$
for this parameter space without subdivision
are given in Table 1.
Errors in the full parameter space would be increased but note that the neglected
parameters $(\Omega_m h^2$,$\Omega_b h^2$, and $n_S$) are exactly those that
the CMB satellite experiments should constrain precisely (see e.g.
\cite{Junetal96} 1996; \cite{EisHuTeg99} 1999).
As an example, we take a sample with $z_{\rm median}=1$ and
$\bar n = 2 \times 10^5$ deg$^{-2}$
as appropriate for a magnitude limit of $R \sim 25$ (see
\cite{Smaetal95b} 1995b).
The signal to noise in the full sample is quite high, e.g.\ at
$\ell=100$, $S/N \sim 25$. Thus we expect that subdividing the sample
should improve parameter estimation.
\begin{center}
{TABLE 1\\[4pt] \scshape Parameter Estimation for $z_{\rm median}=1$} \\[3pt]
\begin{tabular}{llllllll}
}%\vspace{2pt}\tableline\tableline}%\vspace{2pt}
}%\vspace{2pt}\tskip $p_\alpha$ & $\sigma_\alpha f_{\rm sky}^{1/2}$
& \multicolumn{6}{c}{Error Improvement} \\
& 1 &
2(${1 \over 2}$) & 2(${1 \over 4}$) & 2(${1 \over 8}$) & 3(${1 \over 3}$)
& 3(${1 \over 4}$) & 3(${1 \over 8}$) \vphantom{\Big[}\\
}%\vspace{2pt}\tableline}%\vspace{2pt}\tskip
$\Omega_\Lambda$
& 0.040
& 6.5
& 6.9
& 5.7
& 7.2
& 7.7
& 6.9\\
$\Omega_K$
& 0.023
& 2.9
& 3.1
& 2.9
& 3.3
& 3.5
& 3.2\\
$m_\nu$
& 0.044
& 1.7
& 2.0
& 2.1
& 2.1
& 2.2
& 2.2\\
$\ln A$
& 0.064
& 1.7
& 2.0
& 2.0
& 2.1
& 2.2
& 2.1
\end{tabular}
\end{center}
As shown in Table~1, subdividing
this sample in equal halves, denoted as 2(1/2),
improves the errors $\sigma_\alpha$ by a factor of 2 to 7. Since the signal in the lower
redshift bin is smaller than in the higher redshift bin, it suffers comparatively
more from the intrinsic noise variance. One can optimize the binning to correct
for this effect.
Dividing
the sample so as to isolate the upper quarter [2(1/4)] improves
the errors modestly whereas isolating the upper eighth deproves them.
We plot the full range as a function of the fraction of galaxies
in the upper bin in
Fig.~\ref{fig:improvement}. Notice that though the improvement factor
is roughly flat from $0.15-0.5$, it drops rapidly
when noise dominates either the upper or lower fraction. If the signal
were the same in both bins, this would occur at $0.04$ and $0.96$ for
$\ell=100$. The fact that the true improvement is skewed to smaller
upper fractions reflects the fact that the signal increases to higher
redshifts.
\begin{figure}[t]
\centerline{\epsfxsize=3.5truein\epsffile{fig3.eps}}
\caption{Tomographic error improvements on $\Omega_\Lambda$ for
$z_{\rm median}=1$. Upper panel: improvement as
a function of the fraction of galaxies in the upper redshift bin for 2
bins versus 3 bins (same fraction in upper two bins).
Lower panel: redshift corresponding to the upper division.}
\label{fig:improvement}
\end{figure}
Moving to three divisions
makes only a small improvement over two. In Table~1 we give the results
of taking 3 bins with an equal number of galaxies in the upper two bins,
e.g.\ [3(1/4)] represents a division by number of ($1/2$, $1/4$, $1/4$).
In fact the errors for three bins can be higher than those with two
if not chosen wisely.
We conclude that for a redshift distribution
of the form given by equation~(\ref{eqn:distribution})
with $z_{\rm median}=1$, $\alpha=1$ and
$\beta=4$,
crude partitioning suffices to
regain most of the redshift information
in adiabatic CDM models where the change in
the growth rate across the distribution is slow and controlled by a small
number of cosmological parameters
How robust are these conclusions against changes in the distribution
and model? A wider redshift distribution offers greater opportunities for
tomography.
For example, let us widen the distribution by taking
$\beta=2$
in equation~(\ref{eqn:distribution}). Then the gains by simply halving the distribution
are a factor of 9.7 for $\Omega_\Lambda$; going to a 3(1/4) scheme raises
this to 12.
These considerations are also relevant for
deeper surveys.
Consider a survey with $z_{\rm median}=2$ and $\bar n=3.6 \times 10^5$
deg$^{-2}$.
The parameter estimation results are given in Table 2. Not only is the overall
improvement from subdivision larger (up to a factor of 24 for three bins)
but the relative improvements
between parameters also changes. This is because even within the adiabatic
CDM paradigm the importance of the different parameters in determining the growth
of structure depends on redshift.
Perhaps the most important aspect of weak lensing tomography is that
it has the ability to falsify the underlying adiabatic CDM model.
For this reason, it is wise to examine the power spectra from the redshift
bins directly, since these are the observables, rather than jump directly
to modelling the data with parameters under the adiabatic CDM framework.
For example, tomography may show that the component that accelerates
the expansion of the universe is not the cosmological constant
or call into question the
fundamental assumption
that structure forms through the gravitational instability of cold
dark matter.
\begin{center}
{TABLE 2\\[4pt] \scshape Parameter Estimation for $z_{\rm median}=2$}\\[3pt]
\begin{tabular}{lllllll}
}%\vspace{2pt}\tableline\tableline}%\vspace{2pt}
}%\vspace{2pt}\tskip $p_\alpha$ & $\sigma_\alpha f_{\rm sky}^{1/2}$
& \multicolumn{5}{c}{Error Improvement} \\
& 1 & 2(${1 \over 2}$) & 2(${1 \over 4}$) &
2(${1 \over 8}$) &
3(${1\over 4}$) &
3(${1\over 8}$) \vphantom{\Big[}
\\
}%\vspace{2pt}\tskip\tableline}%\vspace{2pt}\tskip
$\Omega_\Lambda$
& 0.063
& 19
& 21
& 20
& 24
& 24\\
$\Omega_K$
& 0.030
& 6.7
& 7.7
& 8.0
& 8.9
& 9.1\\
$m_\nu$
& 0.027
& 2.3
& 2.9
& 3.0
& 3.2
& 3.4 \\
$\ln A$
& 0.040
& 2.1
& 2.6
& 2.1
& 3.1
& 3.2
\end{tabular}
\end{center}
\section{Discussion}
We have shown the precision with which cosmological parameters can
be measured from a weak-lensing survey can be significantly enhanced
by tomographically determining the evolution of the
statistical properties of
large-scale structure across the finite redshift width of the
source distribution. Crude redshift binning of the data can recover
most of the statistical information contained in the redshifts.
For example, most of the gain for a magnitude limited survey
with $z_{\rm median}=1$, under the adiabatic cold dark matter paradigm,
comes from separating out
the upper and lower redshift halves of the distribution.
For wider distributions and stronger rates of change
in the growth of structure, more information can be
extracted by finer binning, especially of the higher redshift portion of the
sample where the signal is greater.
The appropriate number of bins can be empirically determined by
examining the correlation between bins and the noise properties of
the data.
We have been assuming that the individual redshifts of the galaxies
will be known sufficiently precisely to determine the redshift distribution
of the subsamples. Realistically, the redshift information will be limited
by the accuracy of photometric redshift techniques which currently
show errors of $\Delta z \sim 0.1$ (68\% CL) for $0.4 \lesssim z \lesssim
1.4$ (\cite{Hogetal98} 1998).
While statistical errors on the large samples of galaxies considered
above are negligible, systematic errors or biases in the technique may
cause problems. It is beyond the scope of this letter to test these
issues fully. To give some feel for their effect, let us consider
the median redshift $z_{\rm median}$ as an additional parameter with
a prior uncertainty from photometric redshifts of the full individtual
error $0.1$.
Including this uncertainty degrades
the precision in the parameters by $3\%$ in the worst case.
While this effect is negligible, more worrying is a bias that is a function
of redshift, especially in the largely untested regime $1.4 \lesssim z
\lesssim 2$, as that can shift the difference between the power spectra
of the subdivisions. Isolating the few percent of galaxies at $z \gtrsim
2.5$, where the techniques are tested, yields gains that are comparable to
the optimal division (see Fig.~\ref{fig:improvement} lower panel), but the
compactness of such galaxies poses an obstacle for measuring the
lensing distortion from the ground (\cite{Steetal96} 1996).
Despite these caveats, this study shows that tomography with weak lensing is
both possible and would substantially improve the precision with which
we can measure the growth of structure in the universe.
{\it Acknowledgements:}
I would like to thank D.J. Eisenstein, D.W. Hogg, J. Miralda-Escude,
D.N. Spergel, M. Tegmark,
J.A. Tyson, M. White, and D. Wittman for useful conversations.
W.H.\ is supported by the Keck Foundation, a Sloan Fellowship,
and NSF-9513835
|
\section{Physics motivation}
Recent interest in R-parity violating (RPV) SUSY decay modes is motivated by the possible high-$Q^2$ event excess at
HERA~\cite{HERA}. When interpretation of the excess through first-generation leptoquarks was excluded by the D\O~\cite{D0-lq}
and CDF~\cite{CDF-lq} experiments, it was suggested~\cite{HERA-RPV} that such an effect could be explained via the $s$-channel
production of a charm or top squark decaying into the $e + jet$ final state. Both the production and the decay vertices
would thereby violate R-parity. Although more recent data has not confirmed the previous event excess, and despite the combined
analysis showed that the anomalous events reported by the H1 and ZEUS experiments were unlikely to originate from the production
of a single $s$-channel narrow resonance~\cite{Bassler}, interest in RPV signatures has not abatted.
The CDF and D\O\ Collaborations have recently performed searches for RPV SUSY~\cite{CDF-RPV,D0-RPV}, and have set new mass limits
on the RPV SUSY particles. Both experiments focussed their searches on the $\lambda'$ couplings, as motivated by the
high-$Q^2$ HERA event excess. The results of the D\O\ searches are extended to the
Run~2 case and the expected sensitivity to the RPV couplings is discussed.
\section{D\O\ Search for RPV neutralino decays}
The D\O\ search for RPV SUSY considered the case of neutralino LSP which decays into a lepton and two quarks due to a finite
RPV $\lambda'$ coupling (see Fig.~\ref{fig:RPV-decay}). Both the
electron and muon decay channels were considered, corresponding to what commonly referred to as $\lambda'_{1ij}$ and
$\lambda'_{2ij}$ couplings, respectively. The corresponding final states contain either $2e $ or $2\mu $
and at least four accompanying jets. Unlike at HERA, this search
is not sensitive to the value of the RPV coupling, as long as it is large enough so that the neutralino decays within the
D\O\ detector. That corresponds to $\lambda' \geq 10^{-3}$, which gives a lot of room, given current indirect
constraints~\cite{Herbi}.
\begin{figure}[thb]
\centerline{\protect\psfig{figure=rpv_decay.eps,height=1.8in}}
\bigskip
\caption{RPV decay of a neutralino LSP into a lepton and two quarks.}
\label{fig:RPV-decay}
\end{figure}
We assume that the neutralino (LSP) pairs are produced in cascade decays of other supersymmetric particles and use all SUSY pair
production mechanisms when generating signal events.
Signal events were generated within the SUGRA framework with the following values of SUSY parameters: $A_0 = 0$, $\mu < 0$
and $\tan\beta = 2$ (the results are not sensitive to the value of $A_0$ .) Center of mass energy of the colliding beams
was taken to be 2 TeV. {\footnotesize ISAJET}~\cite{ISAJET} was used for event generation. The acceptance and resolution
of the D\O\ detector were parametrized using the following resolutions:
$\delta E/E = 2 \% \oplus 15\%/\sqrt{E}~\mbox{[GeV]}$ (electrons), $\delta (1/p)/(1/p) = 0.018 \oplus 0.008 (1/p)$ (muons), and
$\delta E/E = 3\% \oplus 80\%/\sqrt{E}~\mbox{[GeV]}$ (jets) and found consistent with the full
detector simulation based on {\footnotesize GEANT}~\cite{GEANT}.
\begin{figure}[htb]
\vspace*{-0.3in}
\centerline{\protect\psfig{figure=rpv_points.eps,width=4in}}
\caption{Points in the $(m_0,m_{1/2})$ SUGRA parameter space used to generate RPV events in the $ee + 4$~jets channel.}
\label{fig:RPV-points}
\end{figure}
Figure~\ref{fig:RPV-points} shows the points in the $(m_0,m_{1/2})$ SUGRA parameter space where signal Monte Carlo events
were generated for the electron channel. Similar points were studied for the muon-decay channel.
\section{Selection criteria for the dielectron channel}
A multijet trigger was used for the analysis of Run~1 data. It was found to be nearly 100\% efficient for the typical RPV signal.
Since Run~2 trigger list will include a similar trigger, we assume trigger efficiency of 100\% and do not perform any trigger
simulations for the Run~2.
The following offline selections were used:
\begin{itemize}
\item
At least two good electrons, the leading one with $E_T(e) > 15$ GeV and the other one with $E_T(e) > 10$ GeV;
\item
Rapidity range $\mid \eta \mid \leq 1.1$ (central calorimeter), or $1.5 \leq \mid \eta \mid \leq 2.5$ (end calorimeters) for all
the electrons;
\item
Energy isolation for the electrons: the EM energy in the R=0.2 cone about the center of gravity of the EM cluster, subtracted from
the total energy in R=0.4 cone, should not exceed 15\% of the EM energy in the R-0.2 cone.
\item
At least four jets with $E_T(j) > 15 $ GeV and $\mid \eta \mid < 2.5$;
\item
The dielectron invariant mass ($M_{ee}$) should not be in the Z-mass interval, ie, $ \mid M_{ee} - M_Z \mid > 15$ GeV/$c^2$.
\end{itemize}
In the present analysis we have dropped the requirement on $H_T = \sum E_T(e) + \sum E_T(j)$ , but retained all other offline
criteria that were used in the previous analysis of data from
Run~I~\cite{D0-RPV}.
\section{Selection in the dimuon channel}
The following event selection requirements were used for the muon decay channel:
\begin{itemize}
\item Two muons, the leading one with $p_T >$ 15 GeV, and the other one with $p_T >$ 10 GeV.
\item Rapidity range $|\eta| < 2.3$ for both muons.
\item Energy isolation requirement for both muons, i.e. the calorimeter energy accompanying the muon in a ($\eta$ $\phi$) cone of
0.4 should be consistent with that from a minimum ionising particle.
\item At least four jets with $E_T(j) > 15$ GeV and $|\eta| < 2.5$;
\end{itemize}
\section{Signal efficiencies}
The number of signal events expected can be written as: $\langle N \rangle = \mathcal{L} \cdot \sigma \cdot \epsilon$, where
$\langle N \rangle$ is the expected number of events for luminosity $ \mathcal{L} $, $\sigma$ is the cross-section, and
$\epsilon$ is the overall efficiency. The efficiency $\epsilon$ can be split into three terms:
$\epsilon = {\epsilon}_{\rm trig} \cdot {\epsilon}_{\rm kin} \cdot {\epsilon}_{\rm id}$. Here $\epsilon_{\rm trig}$ is the
trigger efficiency for the events that pass the offline cuts ( assumed to be 100\%), $\epsilon_{\rm kin}$ is the
efficiency for offline criteria, which includes kinematic, fiducial and topological requirements, and
$\epsilon_{\rm id}$ is the electron/jet
identification efficiency.
The efficiency for identifying jets is very high ($> 95\%$) and is expected to stay the same in Run 2.
Electron identification efficiencies in Run 1 were $80 \pm 7\%$ in the central ($|\eta| < 1.1$) and $71 \pm 7\%$ in the forward
($1.5 < |\eta| < 2.5$) regions~\cite{D0-RPV}. These efficiencies were calculated for electrons with $E_T(e) > 25$~GeV, It drops
by about 30\% for electrons with $E_T(e) = 10$~GeV.
The muon identification efficiencies used in Run 1 were $62 \pm 2\%$ in the central ($|\eta| < 1.0$) and $24 \pm 4\%$ in the
forward ($1.0 < |\eta| < 1.7$) regions~\cite{rvmu2}. These were calculated for muons with $p_T > 15$~GeV. For muons with
$10~\mbox{GeV} < p_T < 15$~GeV the efficiencies were 80\% smaller on average~\cite{rvmu3}.
In the present analysis we have taken the overall particle identification efficiency to be $0.90 \pm 0.09$ in each channel,
independent
of lepton $E_T$, primarily due to the expectation of a better tracker and muon spectrometer for the upgraded D\O\ experiment.
\section{Backgrounds}
The main backgrounds are expected to arise from Drell-Yan production in association with
four or more jets, dilepton top-quark events, and QCD multijet events. The latter is
the dominant background for the electron
channel (followed by the Drell-Yan background). In the case of muons, the background is dominated by the Drell-Yan and top pair
production. We used Monte Carlo to calculate background from the first two sources, and data to estimate background from
QCD jets.
Background for the Run 1 analysis was estimated to be $1.8\pm 0.2\pm 0.3$ (with $1.27 \pm 0.24$ from QCD and
$0.42\pm 0.15\pm 0.16$ from the other processes) for $\sim 100\ pb^{-1}$ of data. To extrapolate this number to
the data set from Run 2, we
have simply multiplied it by the ratio of luminosities to obtain $36\pm 4 \pm 6$ events. However, it is expected that due to the
central magnetic field in the upgraded D\O\ detector, the probability of jets to be misidentified as electrons will be reduced by
a factor of $\sim 2$ in Run 2. We have therefore considered a second scenario with the smaller expected background of
$15 \pm 1.5 \pm 1.5$ events.
For the muon channel, the expected background has been scaled directly from the Run 1 analysis. We expect $10 \pm 1 \pm 1$
background events in Run 2.
\section{Results}
In order to obtain the sensitivity of Run 2 in to RPV decays, we calculated the efficiency for signal for
all the mass points shown in Fig.~\ref{fig:RPV-points}. Typical efficiencies,
the signal cross section in the $ee + 4$~jets
channel, and the
expected event yield in 2~$fb^{-1}$ of data, for several representative $(m_0,m_{1/2})$ points, are given in
Table~\ref{table:RPV-eff}. Similar numbers are obtained for the muon channel.
\begin{table}[htb!]
\begin{center}
\caption{Efficiency $\times$ BR (\%), signal cross section and the expected event
yield in 2~$fb^{-1}$ of data, at various $(m_0,m_{1/2})$ parameter space points. }
\begin{tabular}{||c|c|c|c|c||}
\hline
$m_0$ & $m_{1/2}$ & Efficiency $\times$ BR (\%)
& Cross section & $\langle N \rangle$ \\
(GeV) & (GeV) & & (pb) & (in 2 $fb^{-1}$) \\
\hline
60 & 235 & $7.9\pm 1.1 $ & 0.16 & $25.2 \pm 3.4 $ \\
60 & 245 & $8.3\pm 1.1 $ & 0.08 & $12.8 \pm 1.7 $ \\
60 & 255 & $8.3\pm 1.1 $ & 0.06 & $10.5 \pm 1.4 $ \\
100 & 220 & $6.1\pm 0.8 $ & 0.10 & $12.2 \pm 1.7 $ \\
100 & 230 & $7.0\pm 1.0 $ & 0.08 & $11.3 \pm 1.5 $ \\
180 & 240 & $7.0\pm 0.9 $ & 0.05 & $7.1 \pm 1.0 $ \\
320 & 240 & $7.1\pm 0.9 $ & 0.05 & $6.9 \pm 1.0 $ \\
\end{tabular}
\end{center}
\label{table:RPV-eff}
\end{table}
We use these efficiencies to obtain exclusion limits in the $(m_0,m_{1/2})$ plane at 95\% CL, assuming that no excess of
events will be observed above the predicted background. The exclusion contours for the electron and muon channel are shown in
Fig.~\ref{fig:RPV-e} and \ref{fig:RPV-mu}, respectively. Numerical values of the limits are summarized in
Table~\ref{table:RPV-limits}.
\begin{table}[hbt!]
\begin{center}
\caption{Lower limits on the squark and gluino masses from Run 2.}
\begin{tabular}{||c|c|c|c||}
\hline
~ & Lower limit on $m_{\tilde{q}}$ & Lower limit on $m_{\tilde{g}}$ & Limit when
$m_{\tilde{q}} = m_{\tilde{g}}$ \\
~ & ( For any $m_{\tilde{g}}$) & ( For any $m_{\tilde{q}}$ ) & \\\hline
\multicolumn{4}{||c||}{Electrons}\\
\hline
Run 1 & 252 GeV & 232 GeV & 283 GeV \\
Run 2 (Scenario I) & 430 GeV & 490 GeV & 490 GeV \\
Run 2 (Scenario II) & 520 GeV & 575 GeV & 585 GeV \\\hline
\multicolumn{4}{||c||}{Muons}\\\hline
Run 2 & 560 GeV & 640 GeV & 665 GeV \\
\end{tabular}
\end{center}
\label{table:RPV-limits}
\end{table}
\begin{figure}[t]
\centerline{\protect\psfig{figure=rpv_electrons.eps,width=\textwidth}}
\caption{Estimated exclusion contour for Run 2 in the $(m_0,m_{1/2})$ plane for
$tan \beta = 2$, $A_0=0$, $\mu <0$, from the $ee + 4$~jets channel.
Scenario I corresponds to a background of $36\pm 4 \pm 6$ events (direct scaling from Run 1); scenario II uses the background
of $15 \pm 1.5 \pm 1.5$ events (scaling, but with improvements in the detector taken into account).}
\label{fig:RPV-e}
\end{figure}
It's worth mentioning that our analysis provides a conservative estimate of the sensitivity achievable in Run~2, since no formal
optimization of the signal vs. background has been performed. We expect that a formal optimization can improve the sensitivity
in the mass reach by 15--20\%.
\begin{figure}[t]
\vspace*{0.1in}
\centerline{\protect\psfig{figure=rpv_muons.eps,width=\textwidth}}
\caption{Estimated exclusion contour for Run 2 in the $(m_0,m_{1/2})$ plane for
$tan \beta = 2$, $A_0=0$, $\mu <0$, from the $\mu\mu + 4$~jets channel for
background of $10 \pm 1.0 \pm 1.0$ (direct scaling from Run 1).}
\label{fig:RPV-mu}
\vfill
\end{figure}
|
\section{Introduction}
It is often claimed that chiral interaction of two-dimensional fermionic gauge models poses
an obstruction to gauge symmetry. In this paper we clarify several aspects of this question
for different regularizations of the chiral fermionic determinant, including the new Faddeevian
regularization case proposed by Mitra\cite{PM}, under the point of view of the Stone's soldering formalism\cite{ms}.
It is worth mentioning that understanding the properties of 2D fermionic actions is crucial
in several aspects. For instance, the 1-cocycle necessary in recent
discussions on smooth functional bosonization \cite{DNS,DS}, which is just the expression of the 2D
anomaly, is known to be the origin of higher dimensional anomalies through a set of descent equations\cite{DTMP}.
Incidentally, the anomaly phenomenon still defies a complete explanation.
This paper is devoted to analyze and explore the
restrictions that the soldering mechanism \cite{ms,ADW,W,AW} imposes
over the regularization ambiguity of 2D chiral fermionic determinants.
The soldering technique that is dimensionally independent and designed
to work with dual manifestations of some symmetry is well suited to
deal with the chiral character of 2D anomalous gauge theories. Recently
\cite{ABW} a new interpretation for the phenomenon of dynamical mass
generation known as Schwinger mechanism\cite{ls}, has been proposed which explores
the ability of the soldering formalism to embrace interference effects.
In that study the interference of right and left gauged Floreanini-Jackiw
chiral bosons \cite{FJ} was shown to lead to a massive vectorial mode,
for the special case where the Jackiw-Rajaraman (JR) regularization parameter is
$a=1$ \cite{JR}\cite{RR}.
After the discovery that the $\chi QED_{2}$ could be consistently
quantized if the regularization ambiguity were properly taken into account, the investigation
on this subject has received considerable attention and emphasis \cite{many}\cite{AAR}.
The quantization of the model was considered from different points of view,
both canonical and functional and the spectrum and unitarity was analyzed by distinct techniques, including the
gauge invariant Wess-Zumino formulation\cite{WZ}, with results consistent with
Ref.\cite{JR}. Despite this spate of interest, a surprising new
result was reported recently by Mitra \cite{PM} showing that a different
regularization prescription was yet possible, leading to new consequences.
He proposed a new (faddeevian) regularization class, materialized by a unique and
conveniently chosen mass term leading to a canonical description with three constraints.
Recall that in \cite{JR} and \cite{RR}, the Hamiltonian framework was structured in terms of
two classes with two ($a > 1$) and four ($a=1$) second-class constraints respectively. Mitra's work brings a
clear interpretation for the reasons leading the bosonization ambiguity
to fit into three instead of two distinct classes, classified according to the
number of constraints present in the model.
It is the main goal of this paper to study the restrictions posed by the
soldering formalism over this new regularization class.
Since soldering has ruled out the two-constraint class solution of Jackiw-Rajaraman as being able to dynamically generate mass via right-left interference, we are led to ask if the new Faddeevian class of chiral bosons proposed by Mitra do interfere constructively to produce a massive vectorial mode. To find an answer to this question we review, in Section 2,
the procedure of \cite{MM} to obtain the multi-parametric regularization effective action based on the Pauli-Villars regularization proposed in \cite{FS}. This effective action is the point of departure for an
extention of the ambit of Ref.\cite{PM} that
is needed to our purpose in this paper and to be developed in Section 3. The bosonised theory satisfying Faddeev's structure for the constraint algebra is studied in the canonical approach. The mass of the photon scalar field is computed and its dependence on the ambiguity parameter is shown to be tantamount to that in Ref.\cite{JR};
the massless sector however is more constrained than its counterpart in \cite{JR},
corroborating the results of \cite{PM}. The restrictions imposed by the
soldering are worked out in Section 4. We find the striking new result that the interference effects lift the parameter dependence by discriminating the value of the only non ambiguous class. Our results give a clear interpretation for the Schwinger mechanism as a left-right interference phenomenon, as suggested by Jackiw\cite{DTMP}. Our findings are further discussed in the final section.
\section{The Effective Action}
In a gauge invariant theory, free of anomalies, the canonical description reveals a couple of first-class constraints, with the Gauss law $G(x)$ appearing as the secondary constraint for the momentum $\pi_0(x)$ corresponding to the scalar potential $A_0(x)$. In an anomalous gauge theory, on the contrary, gauge invariance is lost and the constraint algebra for the gauge generator becomes afflicted by the presence of a Schwinger term
\begin{eqnarray}
\label{G1}
\left [G(x) , \pi_0(y)\right ] &=& 0\nonumber\\
\left [ G(x) , G(y)\right ] &=& \imath \, \hbar \,{\cal C} \,\delta '(x-y) ,
\end{eqnarray}
where ${\cal C}$ is some constant. This structure introduces extra degrees of freedom into the quantum theory as argued by Faddeev\cite{LF}. The quantum chiral Schwinger model with the usual regularization ($a \geq 1$) does have more degrees of freedom than its classical counterparts, as expected, but does not fit into Faddeev's scheme above due to the functional dependence of the Gauss generator on the scalar potential, which leads to a different constraint algebra than (\ref{G1}),
\begin{eqnarray}
\label{G2}
\left [G(x) , \pi_0(y)\right ] &\not=& 0\nonumber\\
\left [ G(x) , G(y)\right ] &=& 0 .
\end{eqnarray}
The second-class nature of the set is then due to the non-commutative character of the primary and secondary constraints.
The new regularization class for the fermionic determinant proposed by Mitra has the virtue of fitting perfectly into Faddeev's picture. In this section we shall review the computation of the fermionic determinant leading to this new scheme.
Our starting point is the action for fermionic sector of the chiral Schwinger model,
\begin{equation}
S = \int\,d^{2}x\,\,\bar{\psi}(x)\,\left[
i\partial\!\!\!/ - q\,\sqrt{\pi}\,A\!\!\!/(x)\,\left(1 + i\,\gamma_{5}\right)\right]\,
\psi(x)
\label{2.1}
\end{equation}
where $\psi(x)$ is a fermionic field and $A_{\mu}$ is the vector gauge field
in a (1 + 1) dimensional spacetime. From this classical action
we obtain the following effective action\cite{FS}
\begin{equation}
\exp\,i\,S_{eff}^{(0)}[A(x)] =
\int\,{\cal D}\psi(x)\,{\cal D}{\bar\psi}(x)\,
\exp\,i\,S\left[{\bar\psi}(x),\psi(x),A(x)\right].
\label{2.2}
\end{equation}
In a formal level this is a nonlocal action that reads,
\begin{equation}
\exp\,i\,S_{eff}^{(0)}[A(x)] = - q^{2}\,\int\,d^{2}x\,A^{\mu}(x)\,
\left(\eta_{\mu\alpha} + \epsilon_{\mu\alpha}\right)\,
\frac{\partial^{\alpha}\partial^{\beta}}{\partial^{2}}\,
\left(\eta_{\beta\nu} - \epsilon_{\beta\nu}\right)\,
A^{\nu}(x) ,
\label{2.3}
\end{equation}
but there is an ambiguity related to the regularization procedure adopted. Let us discuss the regularization procedure
proposed by Frolov and Slavnov\cite{FS}. To this end we add a multi-parametric
regularising action
\begin{equation}
S_{reg}[A(x)] = \sum_{r=1}^{2n-1}\int\,d^{2}x\,\,
{\bar \psi}_{r}(x)\,\left[ i\,\partial\!\!\!/ - m_{r} - q\,\sqrt{\pi}\,
A_\mu(x)\,\Gamma_{r}^{\mu}\right]\,\psi_{r}(x)
\label{2.4}
\end{equation}
where
\begin{equation}
\Gamma_{r}^{\mu} = \left[a_{r}\,K^{\mu\nu}\,
\left(1 + i\gamma_{5}\right) + b_{r}\,\Sigma^{\mu\nu}\,
\left(1 - i\gamma_{5}\right)\right]\,\gamma_{\nu} .
\label{2.5}
\end{equation}
Here $\psi_{r}(x)$ are the regulators fields with mass $m_{r}$ whose couplings
$\Gamma_{r}^{\mu}$ (or $K^{\mu\nu}$ and $\Sigma^{\mu\nu}$) are matrices
which will be determined later. These regulators bring up the following
partition function
\begin{equation}
\exp\,i\,S_{reg}^{eff}[A] =
\int\,\Pi_{r}\,{\cal D}\psi(x)\,{\cal D}{\bar\psi}(x)\,
\exp\,i\,S_{reg}\left[{\bar\psi}_{r}(x),\psi(x)_{r},A(x)\right]
\label{2.6}
\end{equation}
which can be solved to \cite{MM}
\begin{equation}
S_{reg}^{eff}[A] = -q^{2}\,\frac{\pi}{2}\,\int\,d^{2}x\,
A_{\mu}(x)\,G^{\mu\nu}(x,y)\,A^{\nu}(y)
\label{2.7}
\end{equation}
with
\begin{equation}
G^{\mu\nu}(x,y) = \int\,\frac{d^{2}p}{(2\pi)^{2}}\,{\bar G}^{\mu\nu}(p)\,
\exp\left[-i \cdot p\, (x - y)\right] .
\label{2.8}
\end{equation}
Now ${\bar G}_{\mu\nu}(p)$ is found to be
\[
{\bar G}^{(r)}_{\mu\nu}(p) = \frac{1}{\pi}\,\left\{\left(
a_{r}^{2}\,T^{1}_{\mu\nu\lambda\kappa} + b_{r}^{2}\,T^{1}_{\mu\nu\lambda\kappa}
\right)\,\left[2\,\left(1 + A_{r}\right)\left(\eta^{\lambda\kappa} -
\frac{p^{\lambda}\,p^{\kappa}}{p^{2}}\right) +
A_{r}\,\eta^{\lambda\kappa}\right]+ 2A_{r} a_{r} b_{r} M_{\mu\nu}\right\}
\]
where
\begin{equation}
A_{r} = 1 - \frac{i}{y_{r}}\,\ln(-1) + {\cal O}(y_{r})
\end{equation}
and also
\begin{eqnarray}
T_{\mu\nu\lambda\kappa}^{1} &=& K_{\mu\rho}\,
\left(\delta_{\lambda}^{\rho} + \epsilon_{\kappa}^{\sigma}\right)\,
K_{\nu\sigma}\,
\left(\delta_{\kappa}^{\sigma} + \epsilon_{\kappa}^{\sigma}\right) \nonumber \\
T_{\mu\nu\lambda\kappa}^{2} &=& \Sigma_{\mu\rho}\,
\left(\delta_{\lambda}^{\rho} + \epsilon_{\kappa}^{\sigma}\right)\,
\Sigma_{\nu\sigma}\,
\left(\delta_{\kappa}^{\sigma} + \epsilon_{\kappa}^{\sigma}\right) \nonumber \\
M_{\mu\nu} &=& \left[
K_{\mu\lambda}\,\left(\eta^{\lambda\kappa} - \epsilon^{\lambda\kappa}\right)\,
\Sigma_{\nu\kappa} + \Sigma_{\mu\lambda}\,
\left(\eta^{\lambda\kappa} - \epsilon^{\lambda\kappa}\right)\,
K_{\nu\kappa}\right] \nonumber \\
y_{r}^{2} &=& \frac{p^{2}}{m_{r}^{2}} .
\end{eqnarray}
Imposing the conditions \cite{FS}
\begin{eqnarray}
\sum_r \epsilon_r a_r^2 &=& \sum_r \epsilon_r b_r^2=0\nonumber\\
\sum_r \epsilon_r m_r a_r^2 &=& \sum_r \epsilon_r m_r b_r^2= \sum_r \epsilon_r m_r a_r b_r =0\nonumber\\
2\sum_r \epsilon_r a_r b_r &=& 1
\end{eqnarray}
where $\epsilon_r = (-1)^{r+1}$ is the Grassman parity and
then letting $m_{r} \rightarrow \infty$ we get,
\begin{equation}
S_{reg}^{eff}[A] = q^{2}\,\frac{1}{2}\,\int\,d^{2}x\,A_{\mu}(x)\,
M^{\mu\nu}(x,y)\,A^{\nu}(y).
\label{2.16}
\end{equation}
Jackiw and Rajaraman found a regularized solution with a
diagonal choice for the matrix
\begin{eqnarray}
M^{\mu\nu}\,=\,\pmatrix{a & 0 \cr
0 & a \cr}\, \delta(x - y),
\label{Matrix0}
\end{eqnarray}
with $a\geq 0$, corresponding to the cases with two and four-constraint's classes. The physical content of these cases, as disclosed by them, was found to correspond to an $a$-dependent massive photon field and a massless fermion for the former, while in the later the photon field was absent. Mitra noticed that the alternative choice
\begin{eqnarray}
M^{\mu\nu}\,=\,\pmatrix{1 & -1 \cr
-1 & -3 \cr}\, \delta(x - y),
\label{Matrix1}
\end{eqnarray}
leads to a new class of solutions with three second-class constraints
and found that the physical spectrum of the model contains a chiral
fermion and a photon field with mass $m=4\,q^2$. To work out the soldering
formalism and obtain the interference contribution coming from the chiral
fermions we need to generalize the regularization dependence of the effective
action. This is done in the next section.
\section{Hamiltonian Analysis and Spectrum}
In their seminal work Jackiw and Rajaraman\cite{JR} showed that the $\chi QED_{2}$ could be
consistently quantized by including the bosonization ambiguity parameter satisfying the condition $a \geq 1$ to avoid tachyonic
excitations. Later on, working out the canonical structure of the model,
Rajaraman\cite{RR} showed that the cases $a>1$ and $a=1$ belonged to distinct classes: the $a=1$ case represents the four-constraints class, while the $a>1$ class presents only two constraints. The latter is a continuous one-parameter class, while the former class is non ambiguous containing only one representative. The consequences
of these distinct constraint structures are that the $a>1$ class presents,
besides the massless excitation also a massive scalar excitation
($m^2=\frac{e^2a^2}{a-1}$) that is not found on the other case.
In the canonical approach the
commutator between the primary and the secondary constraints vanishes in the
first case. The emergence of two more constraints
completes the second-class set. Mitra found the amazing fact that with an
appropriated choice of the regularization mass term it is possible to
close the second-class algebra with only three constraints.
His model is not manifestly Lorentz invariant, but the Poincar\'e generators have been constructed \cite{PM} and shown to close the relativistic algebra on-shell.
The main feature of this new regularization is the presence of a Schwinger term in the Poisson bracket algebra of the Gauss law, which limits the set to only three second-class constraints. To see this let us
write the CSM Lagrangian, with faddeevian regularization but with Mitra's
regulator properly generalized to meet our purposes,
\begin{equation}
{\cal L} = -\frac 14 \, F_{\mu\nu}\,F^{\mu\nu} +
\frac{1}{2}\,\partial_\mu\phi\,\partial^{\mu}\phi +
q\,\left(g^{\mu\nu} + b\,\epsilon^{\mu\nu}\right)\,
\partial_{\mu}\phi\,A_{\nu} +
\frac{1}{2}\,q^2\,A_{\mu} M^{\mu\nu} A_{\nu}\,,
\label{Lagrangian}
\end{equation}
where $F_{\mu\nu} = \partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}$; $g_{\mu\nu} =
\mbox{diag}(+1,-1)$ and $\epsilon^{01} = -\epsilon^{10} = \epsilon_{10} = 1$.
$b$ is a chirality parameter, which can assume the values $b=\pm 1$. The
mass-term matrix $M_{\mu\nu}$ is defined as
\begin{eqnarray}
M^{\mu\nu}\,=\,\pmatrix{1 & \alpha \cr
\alpha & \beta \cr}\, \delta(x - y).
\label{Matrix}
\end{eqnarray}
Notice that we have chosen unity coefficient for $A_{0}^{2}$ term. In a sense, this
choice resembles Rajaraman's $a=1$ class and is the trademark of the
faddeevian regularization. In fact, Rajaraman's class is a singular point
in the ``space of parameters''. Its canonical description has the maximum
number of constraints with no massive excitation.
Such a case is found in (\ref{Lagrangian}) if we make $\alpha = 0$ in (\ref{Matrix}). The
appearance of a new class needs a non vanishing value for $\alpha$.
With Mitra's choice, $\alpha = -1$ and $\beta = -3$, the photon becomes massive ($m^2=4 \, q^2$),
but the remaining massless fermion has a
definite chirality, opposite to that entering the interaction with the
electromagnetic field. This choice is, however, too restrictive and may be relaxed
leading to new and interesting consequences. In this work the coefficients
$\alpha$ and $\beta$ are in principle arbitrary, but
the mass spectrum will impose a constraint
between them. This is best seen in the Hamiltonian formalism.
The canonical Hamiltonian is readily computed
\begin{eqnarray}
H = \int\!\!\! &dx& \!\!\! \left\{
\frac{1}{2}\,\left(\pi^{1}\right)^{2} +
\frac{1}{2}\,\pi_{\phi}^{2} +
\pi^{1}A_{0}^{'} +
\frac{1}{2}\,\phi^{'2} + q\,(\,b\phi'\,-\,\pi_{\phi}\,)\,A_0 +
q\left(\phi^{'} - b\,\pi_{\phi}\right)A_{1}
+ \right. \nonumber \\
& & \left. +\,q^{2}\left(b - \alpha\right)A_{0}A_{1} +
\frac{1}{2}q^{2}\left(1-\beta\right)A_{1}^{2}
\right\}\, .
\label{Hamiltonian}
\end{eqnarray}
The stationarity algorithm leads a set of three constraints
\begin{eqnarray}
\label{omega}
\Omega_{1} &=& \pi^{0} , \nonumber \\
\Omega_{2} &=& \left(\pi^{1}\right)^{'} +
q\left(\pi_{\phi}-\,b\,\phi^{'}\right) -
q^{2}\left(b - \alpha\right)A_{1} , \\
\Omega_{3} &=& -\left(b -\alpha\right)\,\pi^{1} +
2\,\alpha\,A_{0}^{'} +
\left(1 +\beta\right)A_{1}^{'} ,\nonumber
\end{eqnarray}
which are easily seen to be second-class, viz.
\begin{eqnarray}
\left\{\Omega_{1}(x), \Omega_{3}(y)\right\} &=& 2\,\alpha\,
\frac{\partial}{\partial x}\,\delta (x-y)\, , \nonumber \\
\left\{\Omega_{2}(x), \Omega_{2}(y)\right\} &=& -2\,q^2\,\alpha\,
\frac{\partial}{\partial x}\,\delta (x-y)\, , \nonumber \\
\left\{\Omega_{2}(x), \Omega_{3}(y)\right\} &=&
q^{2}\left(b-\alpha\right)^{2}\,\delta (x-y)-
\left(1+\beta\right)\,\frac{\partial^{2}}{\partial x \, \partial y}\,
\delta (x-y)\, , \nonumber \\
\left\{\Omega_{3}(x), \Omega_{3}(y)\right\} &=& -2\left(b-\alpha\right)
\left(1+\beta\right)\frac{\partial}{\partial x}\,\delta (x-y)\,\,,
\end{eqnarray}
with the other brackets vanishing. This is in sharp contrast with the usual regularization possessing two or four second-class constraints. To perform quantization we compute the Dirac
brackets,
\begin{eqnarray}
\label{db}
\left\{\,\phi(x)\,,\,\phi(y)\,\right\}_{D} &=& -\frac{1}{4\,\alpha}\,
\theta (x-y)\, , \nonumber \\
\left\{\phi(x),A_{1}(y)\right\}_{D} &=& -\frac{1}{2\,q\,\alpha}\,
\delta (x-y)\, , \nonumber \\
\left\{\phi(x),\pi^{1}(y)\right\}_{D} &=& -\frac{q}{4\,\alpha}
\left(b - \alpha\right)\,
\theta (x-y)\, , \nonumber \\
\left\{A_{1}(x),A_{1}(y)\right\}_{D} &=& \frac{1}{2\,q^{2}\,\alpha}\,
\frac{\partial}{\partial x}\,
\delta (x-y)\, , \\
\left\{\pi^{1}(x),A_{1}(y)\right\}_{D} &=& -\left(\frac{b+\alpha}{2\,\alpha}
\right)\delta (x-y)\, ,
\nonumber \\
\left\{\pi^{1}(x),\pi^{1}(y)\right\}_{D} &=& -\frac{q^{2}}{4\,\alpha}
\left(b-\alpha\right)^{2}
\theta (x-y)\,\,. \nonumber
\end{eqnarray}
The reduced Hamiltonian is obtained by strongly eliminating $\pi^{0}$, $A_{0}^{'}$
and $\pi_{\phi}$ from the constraints (\ref{omega}) and substituting in the canonical
Hamiltonian (\ref{Hamiltonian}),
\begin{eqnarray}
\label{hr}
H_{r} = \int\!\!\! &dx& \!\!\! \left\{
\frac{1}{2}\left(\pi^{1}\right)^{2} -
\alpha\,\pi^{1}A_{1}^{'} + q\left(1-b\,\alpha\right)A_1\phi^{'} +
\phi^{'2}-\frac{b}{q}\,\phi^{'}
\left(\pi^{1}\right)^{'} +
\right. \nonumber \\
&& \left. + \frac{1}{2\,q^{2}}
\left(\pi^{1}\right)^{'2} +\frac{1}{2}\,q^{2}\left(\alpha^{2}-
\beta\right)A_{1}^{2}\right\}\, .
\end{eqnarray}
Making use of (\ref{db}) and (\ref{hr}) we get the following equations of
motion for the remaining fields,
\begin{eqnarray}
\dot{\phi} &=& b\,\phi^{'} - \frac{1}{q}
\left(\pi^{1}\right)^{'} + \frac{q}{2\,\alpha}
\left(1 - 2\alpha^{2}+\beta\right)A_{1} \label{phi_eq}\, ,\label{phi} \\
\dot{\pi}^{1} &=& -b
\left(\pi^{1}\right)^{'} +\frac{q^{2}}{2\,\alpha}\left[(b-\alpha)
(1-\alpha^{2}) - (b+\alpha)(\alpha^{2}-\beta)\right]A_{1} \label{pi1}\, , \\
\dot{A}_{1} &=& \left(\frac{\alpha + b}{2\alpha}\right)\pi^{1} -
\left(\frac{1 + \beta}{2\alpha}\right)A_{1}^{'}\,\, . \label{a1}
\end{eqnarray}
We are now ready to determine the spectrum of the model. Isolating $\pi^{1}$
from the Eq.(\ref{a1}) and substituting
in the equation (\ref{pi1}), we will have
\begin{eqnarray}
\left(\frac{2\,\alpha}{\alpha + b}\right)\ddot{A}_{1}+
b\left(\frac{1 + \beta}{\alpha + b}\right)A_{1}^{''}\!\!\! &=& \!\!\!
-\left(\frac{2\,b\,\alpha}{\alpha + b} + \frac{1+\beta}{\alpha +b}\right)
\dot{A}_{1}^{'}\,+\, \nonumber \\
\!\!\! & + & \!\!\! \frac{q^2}{2\alpha}\left[\left(b - \alpha\right)
\left(1-\alpha^{2}\right) - \left(b + \alpha\right)
\left(\alpha^{2}-\beta\right)\right]A_{1}\, .
\label{eq}
\end{eqnarray}
To get a massive Klein-Gordon equation for the photon field
we must set
\begin{equation}
\label{condicao}
\left(1 + \beta\right) + b\,\left(2\alpha\right) = 0\, ,
\end{equation}
which relates $\alpha$ and $\beta$ and shows that the regularization ambiguity adopted in
\cite{PM} can be extended to a continuous one-parameter class (for a chosen
chirality). We have, using (\ref{eq}) and (\ref{condicao}), the following mass
formula for the massive excitation of the spectrum,
\begin{equation}
m^{2} = q^{2}\,\frac{\left(1 + b\,\alpha\right)^{2}}{b\,\alpha}\, .
\end{equation}
Note that to avoid tachyonic excitations, $\alpha$ is further restricted to
satisfy $b\,\alpha = |\alpha|$, so $\alpha \rightarrow -\alpha$
interchanges from one chirality to another.
Observe that in the limit $\alpha \rightarrow 0$ the massive
excitation becomes infinitely heavy and decouples from the spectrum. This
leads us back to the four-constraints class. It is interesting to see that the redefinition of
the parameter as $a=1+ |\alpha|$ leads to,
\begin{equation}
m^2 = \frac{q^2 a^2}{a-1}
\end{equation}
which is the celebrate mass formula of the chiral Schwinger model, showing that the parameter dependence of the mass spectrum
is tantamount to both the Jackiw-Rajaraman and the faddeevian
regularizations.
Let us next discuss the massless sector of the spectrum. To disclose the
presence of the chiral excitation we need to diagonalize the reduced
Hamiltonian (\ref{hr}). This procedure may, at least in principle, impose further restrictions over $\alpha$.
This all boils down to find the correct linear
combination of the fields leading to the free chiral equation of motion. To this end
we substitute $\pi^{1}$ from its definition and $A_{1}$ from
the Klein-Gordon equation into equation (\ref{phi_eq}) to obtain
\begin{eqnarray}
\label{vinte}
0 &=&\frac{\partial}{\partial t}\left\{
\phi + \frac{q}{2\alpha}
\left(\frac{2 + 2\,b\,\alpha - \alpha^{2}}{m^{2}}\right)
\dot{A}_{1} + \frac{1}{q}\left(\frac{\alpha}{\alpha + b}
\right)A_{1}^{'}\right\}\,-\, \nonumber \\
&-& \frac{\partial}{\partial x}\left\{
b\,\phi - \frac{1}{q}\left(\frac{\alpha}{\alpha + b}\right)\dot{A}_{1} +
\left[\frac{q}{2\alpha}\left(\frac{2 + 2\,b\,\alpha -\alpha^{2}}{m^{2}}\right)
-\frac{1}{q}\left(\frac{2\,b\,\alpha}{\alpha+b}\right)\right]\right\}\,\,.
\end{eqnarray}
This expression becomes the equation of motion for a self-dual boson $\chi$
\begin{equation}
\dot{\chi} - b\,\chi^{'} = 0
\label{quiral}
\end{equation}
if we identify the coefficients for $\dot A_1$ and $A_1^{'}$ in the two independent terms of (\ref{vinte}) with,
\begin{equation}
\label{chi}
\chi = \phi + \frac 1{q} \left(\frac{\alpha}{\alpha + b}\right)\left(A_{1}^{'}- b \dot{A}_{1}\right).
\end{equation}
This field redefinition, differently from the case of the massive field whose construction imposed condition (\ref{condicao}), does not restrain the parameter $\alpha$ any further.
Using
(\ref{omega}) and (\ref{quiral}), all the fields can be expressed as
functions of the free massive scalar $A_1$ and the free chiral boson
$\chi$, interpreted as the bosonized Weyl fermion. The main result of this section is now complete, i.e., the construction of the one-parameter class regularization generalizing Mitra's proposal. The stage is now set to study the interference of chiral actions with (one-parameter) faddeevian regularization.
\section{Effects of Interference}
In this section we use the soldering formalism introduced in
\cite{ms} to examine the restriction imposed by chiral interference over
the regularization ambiguity parameter when the faddeevian approach is adopted.
This study, taken in the framework of the usual JR regularization, establishes a strong restriction over the parameter's values and gives rise to a new interpretation
for the mechanism of dynamical mass generation occurring in the Schwinger model.
This study is meaningful and necessary since a new class of theories with three
second-class constraints has emerged: it must be verified if new solutions resulting from interference will lead to a gauge invariant massive excitation. To begin with, let
us rewrite explicitly the two chiral actions presented in (\ref{Lagrangian})
in the appropriate light-cone variables,
\begin{eqnarray}
{\cal L}_+\,&=& \partial_{+}\rho\,\partial_{-}\rho + \frac{1}{2}\,
\left(\partial_{-}A_{+} - \partial_{+}A_{-}\right)^{2} +
2\,q\,\partial_{-}\rho\,A_{+}
-2\, q^{2}\,|\alpha|\,A_{-}^{2} +\nonumber \\
& &+\;q^{2}\,\left(1\,+\,|\alpha|\right)A_{-}A_{+} \label{mitra+}\\
{\cal L}_-\,&=&\partial_{+}\varphi\,\partial_{-}\varphi + \frac{1}{2}\,
\left(\partial_{-}A_{+} - \partial_{+}A_{-}\right)^{2} +
2\,q\,\partial_{+}\varphi\,A_{-}
-2\, q^{2}\,|\bar\alpha|\,A_{+}^{2} +\nonumber \\
& &+\;q^{2}\,\left(1\,+\,|\bar\alpha|\right)A_{-}A_{+} \;\;,
\label{mitra-}
\end{eqnarray}
where we have used the convention ${\cal L}_{\pm} = {\cal L}|_{b}$. For clarity, we have used different fields $(\varphi, \rho)$ for opposite chiralities and the corresponding
mass-term parameters $(\alpha,\overline{\alpha})$ to make clear that these
chiral theories are uncorrelated. However by making use of soldering
formalism we will get a meaningful combination of these components.
The main point of soldering is to lift the global Nother symmetry of each chiral component to a local symmetry of the system as a whole. Showing only the main parts of the soldering
formalism we can see that the axial transformation ($\delta\varphi =\delta\rho = \eta$) leads to
\begin{eqnarray}
\label{CW10}
\delta {\cal L}_+\,&=&\, \partial_-\eta\,J_+(\rho) \nonumber \\
\delta {\cal L}_-\,&=&\, \partial_+\eta\,J_-(\varphi)
\end{eqnarray}
where $J_{-}(\varphi)=2(\partial_-\varphi\,+\,q A_-)$ and $J_{+}(\rho)=2(\partial_+\rho\,+\,q A_+)$ are the Noether's currents and $\eta$ is the gauge parameter. Next we introduce the soldering field
$B_{\pm}$ appropriately coupled to the Noether currents to obtain the once iterated chiral actions as,
\begin{eqnarray}
{\cal L}_+^{(0)} &\rightarrow& {\cal L}_+^{(1)}\,=\,{\cal L}_+^{(0)}\,+\,B_+\,
J_-(\varphi) \nonumber \\
{\cal L}_-^{(0)} &\rightarrow& {\cal L}_-^{(1)}\,=\,{\cal L}_-^{(0)}\,+\,B_-\,
J_+(\rho) \;\; .
\end{eqnarray}
The soldering fields act as partial compensators for the variance (\ref{CW10}), transforming vectorially under the axial symmetry, $\delta B_{\pm}=\partial_{\pm}\eta$.
It is now possible to define an effective Lagrangian
invariant under the combined transformation of the chiral fields and compensators as,
\begin{equation}
{\cal L}_{eff}\,=\,{\cal L}_+^1\,+\,{\cal L}_-^1\,+\,2\,B_+\,B_-\;\;.
\label{Leff}
\end{equation}
The soldered action is obtained using the fact that $B_{\pm}$ are auxiliary fields. Their elimination may be done altogether from their field equations but the effects of
soldering will persist as a residual symmetry for the remaining fields. This will naturally cohere the otherwise
independent chiral fields $\varphi$ and $\rho$ in the form of a
soldered Lagrangian for a collective field $\Phi$ as,
\begin{eqnarray}
\label{solda}
{\cal L}_{eff}\,&=&\,\partial_+\,\Phi\,\partial_-\,\Phi\,-\,2\,q\,
(A_+\,\partial_-\,
\Phi\,-\,A_-\,\partial_+\,\Phi)\,+\, \frac 12 \left(\partial_+ A_-\, -\, \partial_- A_+\right)^2\nonumber\\
\,&+&\,{q^2 \over 2}\left[\alpha\,A_+^2\, - \,\overline{\alpha}\,A_-^2\, - \,\left(\alpha - \,\overline{\alpha}\right) A_+\,A_-\,\right]
\end{eqnarray}
where $\Phi=\varphi-\rho$. Notice that except for the last term, the soldered action describes the massive gauge invariant bosonised version of the Schwinger model, with the gauge invariant collective field $\Phi$ playing the role the of the photon field. Gauge invariance imposes a strong constraint over the parameters as,
\begin{equation}
\alpha\,= \,\overline{\alpha} = 0.
\end{equation}
This value corresponds to the $a=1$, four-constraints regularization class. This is a remarkable result, consistent with \cite{ABW}.
A notable feature of the present analysis is the disclosure of a new class of
parameterizations and their dependence with the number of constraints.
Different aspects of this feature were
elaborated and the consequences of interference computed.
To discuss further the implications of interference on chiral actions it is best to compare with the existing literature. This also serves
to put the present work in a proper perspective.
To be precise, it was initialy shown that in the faddeevian approach there are actually three
second-class constraints with a real parameter dependence.
To disclose this one-parameter dependence of the faddeevian regularization is a new interesting result. The counting of constraints explains the presence of only one chiral excitation in the spectrum (besides the massive mode).
This is in contrast with the usual JR regularization where the massless excitation is scalar, and is essentially tied to the fact that this regularization is less constrained.
The restrictions of soldering however confine the appearance of a massive vector excitation to the interference of modes belonging to the $a=1$ class that, being more constrained, has only a massless scalar in the spectrum.
This might raise questions about the interference of the chiral modes in this class.
It should be noticed however that the use of light-cone variables in the soldering constrains even further these chiral actions. Both the two and the four-constraints classes display chiral excitations instead of massless scalars. The original chiral mode of the Mitra's class therefore disappear in the presence of the extra light-cone constraint and there appears to exist an ambiguity challenging the real meaning of the soldering. In fact there are no massless particles in the spectrum
of (\ref{mitra+}-\ref{mitra-}) for the light-cone setting.
However, what is important to observe in this scenario is that the whole process of soldering is done in the Lagrangian framework, such that the limit $a\rightarrow 1$ is well defined. This is also valid for the JR regularization. The limit leads to the $a=1$ action and the canonical analysis may be done unambiguously. Oppositely, the Hamiltonian formulation has the $a=1$ point as a singularity, as shown in (\ref{db}).
\section{Conclusions}
In this work we studied the bosonized form of the CSM fermionic determinant adopting the three-constraints regularization parameterized by a single real number. This extends early regularizations
proposed by JR and Mitra. Our results display a clear-cut separation of the
existing classes shown to depend only on the number of
second-class constraints. The new class with faddeevian regularization and three second-class constraints has been worked out in great detail. The spectrum
has been shown to consist of a chiral boson and a massive photon field.
The mass formula for the scalar excitation was shown to reproduce the JR result. Considerations of unitarity therefore restrain the range of the regularization parameter similarly. The use of the
soldering formalism supplemented by gauge invariance restricts the otherwise arbitrary ambiguity parameter to the specific value $a=1$, which corresponds to the four-constraints class. This is new result that discriminates the special character of this unambiguous regularization point and gives a precise interpretation of the Schwinger dynamical mass generation mechanism as a consequence of right and left interference.
To conclude we stress that the formalism and analysis proposed here illuminates the close
connection among anomalous gauge theories, the interference phenomenon and the mechanism
of dynamical mass generation, providing a variety of new possibilities with practical
applications.
\bigskip
\noindent {\bf Acknowledgment:} This work is supported in part by
CNPq, FINEP, CAPES, FAPESP and FUJB (Brazilian Research Agencies).
|
\section{#1}\setcounter{equation}{0}}
\def\baselinestretch{1.1}\textheight 23.5cm\textwidth 16cm\parskip 1ex
\oddsidemargin 0pt\evensidemargin 0pt\topmargin -40pt\jot = .5ex
\setlength{\parskip}{2mm}
\newcommand{\tilde{x}_t}{\tilde{x}_t}
\newcommand{\rightarrow}{\rightarrow}
\newcommand{\mu}{\tilde{M}_p}
\newcommand{\tilde{g}_s}{\tilde{g}_s}
\newcommand{\tilde{v}_t}{\tilde{v}_t}
\newcommand{\tilde{M}_s}{\tilde{M}_s}
\newcommand{\begin{equation}}{\begin{equation}}
\newcommand{\end{equation}}{\end{equation}}
\newcommand{\eel}[1]{\label{#1}\end{equation}}
\newcommand{\begin{eqnarray}}{\begin{eqnarray}}
\newcommand{\end{eqnarray}}{\end{eqnarray}}
\newcommand{\eeal}[1]{\label{#1}\end{eqnarray}}
\newcommand{\begin{equation}\begin{array}{rcl}}{\begin{equation}\begin{array}{rcl}}
\newcommand{\end{array}\end{equation}}{\end{array}\end{equation}}
\newcommand{\eaql}[1]{\end{array}\label{#1}\end{equation}}
\newcommand{\begin{equation}\begin{array}{rcl}}{\begin{equation}\begin{array}{rcl}}
\newcommand{\eeacn}[1]{\end{array}\label{#1}\end{equation}}
\newcommand{\begin{array}}{\begin{array}}
\newcommand{\end{array}}{\end{array}}
\newcommand{\nonumber \\}{\nonumber \\}
\newcommand{\equ}[1]{(\ref{#1})}
\renewcommand{\a}{\alpha}
\renewcommand{\b}{\beta}
\renewcommand{\d}{\delta} \newcommand{\Delta}{\Delta}
\newcommand{\varepsilon}{\varepsilon}
\newcommand{\lambda} \newcommand{\La}{\Lambda}{\lambda} \newcommand{\La}{\Lambda}
\newcommand{\nu}{\nu}
\newcommand{\omega} \newcommand{\OM}{\Omega}{\omega} \newcommand{\OM}{\Omega}
\newcommand{\psi} \newcommand{\PS}{\Psi}{\psi} \newcommand{\PS}{\Psi}
\newcommand{\s}{\sigma}
\newcommand{{\phi}} \newcommand{\F}{{\Phi}}{{\phi}} \newcommand{\F}{{\Phi}}
\newcommand{{\varphi}}{{\varphi}}
\newcommand{{\upsilon}} \newcommand{\Y}{{\Upsilon}}{{\upsilon}} \newcommand{\Y}{{\Upsilon}}
\newcommand{\zeta}{\zeta}
\renewcommand{\AA}{{\mathcal A}}
\newcommand{{\mathcal B}}{{\mathcal B}}
\newcommand{{\mathcal C}}{{\mathcal C}}
\newcommand{{\mathcal D}}{{\mathcal D}}
\newcommand{\bigtriangledown}{\bigtriangledown}
\newcommand{{\alpha^{'}}}{{\alpha^{'}}}
\newcommand{\begin{eqnarray}}{\begin{eqnarray}}
\newcommand{\end{eqnarray}}{\end{eqnarray}}
\newcommand{\nonumber}{\nonumber}
\newcommand{\,\raisebox{-1.0ex}{\,\raisebox{-1.0ex}
{$\stackrel{\textstyle <}{\sim}$}\,}
\newcommand{\,\raisebox{-1.0ex}{$\stackrel{\textstyle >}{\,\raisebox{-1.0ex}{$\stackrel{\textstyle >}
{\sim}$}\,}
\newcommand{Schwarzschild $\:$}{Schwarzschild $\:$}
\newcommand{\varphi}{\varphi}
\newcommand{\Delta}{\Delta}
\newcommand{\tilde{x}}{\tilde{x}}
\newcommand{\hspace{-.65cm}}{\hspace{-.65cm}}
\newcommand{\journal}[4]{{\rm #1~}{#2}\,(19#3)\,#4}
\newcommand{\journal {Ann. Inst. Henri Poincar\'e}}{\journal {Ann. Inst. Henri Poincar\'e}}
\newcommand{\journal {Helv. Phys. Acta}}{\journal {Helv. Phys. Acta}}
\newcommand{\journal {Sov. J. Part. Nucl.}}{\journal {Sov. J. Part. Nucl.}}
\newcommand{\journal {Int. J. Mod. Phys.}}{\journal {Int. J. Mod. Phys.}}
\newcommand{\journal {Physica (Utrecht)}}{\journal {Physica (Utrecht)}}
\newcommand{\journal {Phys. Rev.}}{\journal {Phys. Rev.}}
\newcommand{\journal {JETP Lett.}}{\journal {JETP Lett.}}
\newcommand{\journal {Phys. Rev. Lett.}}{\journal {Phys. Rev. Lett.}}
\newcommand{\journal {J. Math. Phys.}}{\journal {J. Math. Phys.}}
\newcommand{\journal {Rev. Mod. Phys.}}{\journal {Rev. Mod. Phys.}}
\newcommand{\journal {J. Diff. Geom.}}{\journal {J. Diff. Geom.}}
\newcommand{\journal {Comm. Math. Phys.}}{\journal {Comm. Math. Phys.}}
\newcommand{\journal {Class. Quantum Grav.}}{\journal {Class. Quantum Grav.}}
\newcommand{\journal {Z. Phys.}}{\journal {Z. Phys.}}
\newcommand{\journal {Nucl. Phys.}}{\journal {Nucl. Phys.}}
\newcommand{\journal {Phys. Lett.}}{\journal {Phys. Lett.}}
\newcommand{\journal {Mod. Phys. Lett.}}{\journal {Mod. Phys. Lett.}}
\newcommand{\journal {Phys. Reports}}{\journal {Phys. Reports}}
\newcommand{\journal {Progr. Theor. Phys.}}{\journal {Progr. Theor. Phys.}}
\newcommand{\journal {Nuovo Cim.}}{\journal {Nuovo Cim.}}
\newcommand{\journal {Acta Phys. Pol.}}{\journal {Acta Phys. Pol.}}
\newcommand{\journal {Astrophys. Jour.}}{\journal {Astrophys. Jour.}}
\newcommand{\journal {Astrophys. Jour. Lett.}}{\journal {Astrophys. Jour. Lett.}}
\newcommand{\journal {Ann. Phys. }}{\journal {Ann. Phys. }}
\newcommand{{\em Phys. Rev. D}}{{\em Phys. Rev. D}}
\newcommand{\journal {Gen. Rel. Grav.}}{\journal {Gen. Rel. Grav.}}
\newcommand{g_{YM}}{g_{YM}}
\newcommand{g_{eff}}{g_{eff}}
\begin{document}
\newcommand{\preprint}[1]{\begin{table}[t]
\begin{flushright}
\begin{large}{#1}\end{large}
\end{flushright}
\end{table}}
\preprint{hep-th/9904035}
\begin{center}
\Large{\bf A Comment on the Entropy of Strongly\\ Coupled ${\cal N}=4$}
\vspace{5mm}
\normalsize{N. Itzhaki}
\vspace{5mm}
{ Department of Physics\\
University of California, Santa Barbara, CA 93106}\\
{\it <EMAIL>}
\end{center}
\begin{abstract}
We propose a field theory argument, which rests on the non-renormalization of
the two point function of the energy-momentum tensor, why the ratio between the entropies of
strongly coupled and weakly coupled ${\cal N}=4$ is of order one.
\end{abstract}
\baselineskip 18pt
The Maldacena conjecture \cite{mal} and the entropy of near-extremal D3-branes
\cite{gkp} imply that the ratio between the entropies, at fixed
temperature, of strongly coupled and weakly coupled ${\cal N}=4$ is $3/4$.
In ${\cal N}=4$, unlike 2D CFT, the entropy is not protected thus it is not surprising that the
ratio is not $1$.
It is surprising, however, that the ratio is not a function of the 't Hooft
coupling, $\lambda= g_{YM}^2 N$, which vanishes when $\lambda \rightarrow \infty$.
The reason is the following {\em perturbative} argument.\footnote{Though
this argument is widely known we did not find
it in the literature. A closely related discussion can be found in
\cite{hms,bjsv}.}
At finite temperature, $T$, the expectation value of the fields is
$\langle \phi^2 \rangle =T^2.$
As a result the potential term in SYM, which has the form
$ V \sim g_{YM}^2 [\phi_i, \phi_j ]^2,$
induces a mass, $m^2 \sim \lambda T^2$, for a generic field.
At small 't Hooft coupling the induced masses are much smaller then the
temperature so to a good approximation the contribution to the entropy is of
$N^2$ massless fields with a small correction
which reduces the entropy.\footnote{For a rigorous discussion
on the weakly coupled region see \cite{ft}.}
At large 't Hooft coupling the induced masses are much larger then the
temperature.
Therefore, the contribution to the entropy from a generic field (not
in the Cartan subalgebra of $SU(N)$) is suppressed at the strongly coupled
region.
Since the argument above rests on perturbation theory it cannot be
trusted all the way to the strongly coupled region and hence, strictly
speaking, there is no contradiction with the Maldacena conjecture.
Still, it is fair to say that it is somewhat disturbing that the only
field theory argument available (as far as we know) leads to that conclusion.
Especially, when a similar argument for SYM in $1+1$ dimensions
\cite{hms,bjsv} leads to results that fit so nicely into the Maldacena
conjecture for D1-branes \cite{juan,pp,aki}.
The purpose of this short note is to put forward a {\em field theory}
argument, which does not rest on the AdS/CFT correspondence,
that implies that the entropy at large coupling is of the order of the
entropy at weak coupling.
The argument rests on the ${\cal N}=4$ non-renormalization theorem for the
energy-momentum tensor two point function and therefore it cannot be generalised
to two dimensional SYM which is a non-conformal theory and hence the R-symmetry
cannot protect the two-point function.
We study SYM in a box whose volume is
$L_z A$ with $A=L_x L_y$ and we take the limit $L_x, L_y \gg L_z$.
Consider the transformation $x_3 \rightarrow x_3(1-\varepsilon)$ with $\varepsilon \ll 1$.
The variation of the action under this transformation is
\begin{equation}
\d S=\varepsilon\int d^3 x \int_0^{L_z} dx_3 T_{33}.
\end{equation}
Therefore, the variation of the expectation value of $T_{00}$ is
\begin{eqnarray}
\d \langle T_{00}(0)\rangle =\int {\cal D} \phi ( e^{-(S+\d S)} -e^{-S})
T_{00}(0)=\varepsilon\int d^3x
\int_{0}^{L_z}dx_3
\langle T_{00}(0) T_{33}(x)\rangle,
\end{eqnarray}
where ${\cal D} \phi$ represents integration with respect to all fields.
To calculate the integral we need to know the energy-momentum tensor two
point function.
On $R^4$ non-renormalization theorem protects
the energy-momentum tensor two point function.
Thus on $R^4$ we can use the free SYM result
\begin{equation}\label{11}
\langle T(0)T(x) \rangle =\frac{N^2}{x^8},
\end{equation}
where we have suppressed numerical factors of order one and the Lorentz
indices (for details see \cite{2p}).
However, what we need is not the two-point function in $R^4$ but
rather in $R^3 \times S^1$.
In two dimensions the conformal transformation group contains the
transformation from $R^2$ to $R \times S^1$.
Therefore, the two-points function in $R \times S^1$ are determined by the
two-points function in $R^2$ and the dimensions of the operators.
This is an important ingredient in Cardy's proof that the asymptotic
growth of the number of state of a 2D CFT depends only on the central charge
and not on the details of the CFT \cite{car,car1}.
In four dimensions, however, the conformal transformations do not
contain the transformation from $R^4$ to $R^3 \times S^1$.
Thus, we do not know the exact form of the energy-momentum two point function
for strongly coupled SYM on $R^3 \times S^1$.\footnote{For weakly
coupled theories one can find directly on $R^3\times S^1$
the mode expansion of the relevant fields.
So there is no need to start with the two-points function on $R^4$.}
What we do know is that at distances smaller then $L_z$
the boundary condition is irrelevant and so
eq.(\ref{11}) is a good approximation at short distances.
Therefore, for a given point on $S^1$ we can calculate the contribution to
$\d \langle T_{00}(x_3)\rangle$ from the region $|x_3-x_3^{'}| < L_z$.
The integral divergent at short distances. The regularized result is
$ \d \langle T_{00}(0)\rangle \sim \frac{\varepsilon N^2}{ L_z^4}$.
Integrating over the volume we find the variation of the ground
state energy, which yields after
integration with respect to $\d L_z=\varepsilon L_z$, the ground state energy
\begin{equation}\label{12}
E_0\sim \frac{N^2 A}{L_z^3}.
\end{equation}
It is important to emphasis that we have assumed in the calculation of
$\d \langle T_{00}\rangle$ that the integration
over the whole region does not contain cancellations between the region where
eq.(\ref{11}) is a good approximation
and the region where it is not.
Such cancellations can, in principle, reduce the ground state energy in a
significant way to yield $E_0$ which is suppressed at large coupling.
Therefore, our argument is not a proof but rather a strong indication
that the entropies ratio is of order one.
In other words, we estimate the Casimir energy, which is a boundary
condition effect, using an approximation which is not sensitive to the details
of the boundary condition but only to the distance between the
boundaries.
Eq.(\ref{12}) implies that the partition function at low temperature
(compared to $L_z$) is,
\begin{equation}
Z\sim \exp \left( \frac{N^2 A \beta}{L_z^3}\right) .
\end{equation}
Now we can use the standard argument of switching the roles of
$\beta $ and $L_z$ \footnote{See \cite{hm} for a related discussion in the context of
the AdS/CFT correspondence.} to end up with the partition function
of strongly coupled SYM at high temperature (compared to the size of the box)
\begin{equation}
Z\sim \exp (N^2 V T^3),
\end{equation}
which agrees, up to a numerical factor in the exponent, with the
partition function of weakly coupled SYM.
\vspace{.5cm}
\hspace{-.65cm} {\bf Acknowledgements}
I would like to thank Aki Hashimoto for helpful discussions.
Work supported in part by the NSF grant PHY97-22022.
|
\section{Introduction}
\label{sec:intro}
Error-correction is required whenever information has to be reliably
transmitted through a noisy environment. The theoretical
grounds for classical error-correcting codes were first presented in
1948 by Shannon \cite{shannon}. He showed that it is possible to transmit
information trough a noisy channel with a vanishing error probability
by encoding
up to a given critical rate $R_c$ equivalent to the
{\it channel capacity}.
However, Shannon's arguments were non-constructive and devising such
codes turned out to be a major practical problem in the area of information
transmission.
In 1989 Sourlas \cite{sourlas89,sourlas94} proposed that, due to the
equivalence between addition over the field $\{0,1\}$ and multiplication over
$\{{\pm 1}\}$, many error-correcting codes can be mapped onto many-body
spin-glasses with appropriately defined couplings. This observation opened
the possibility of applying techniques from statistical physics to
study coding systems, in particular, these ideas were applied to the study of
parity check codes.
These linear block codes can be represented by matrices of $N$ columns
and $M$ rows that transform $N$-bit messages to $M$ ($>N$) parity checks.
Each row represents bits involved in a
particular check and each column represents checks involving
the particular bit. The number of bits used in each check and the number
of checks per bit depends on the code construction. We concentrate on the
case where exactly $C$ checks are performed for each bit and exactly
$K$ bits compose each check.
The {\it code rate} $R$ is defined as the information conveyed per channel
use $R=H_2(f_s)N/M=H_2(f_s)K/C$, where $H_2(f_s)=
-(1-f_s)\;\mbox {log}_2 (1-f_s)\;-
\;f_s\;\mbox {log}_2 (f_s) $ is the binary entropy of the message with bias
$f_s$.
\begin{figure}
\hspace*{.4cm}
\epsfxsize=150mm \epsfbox{figure1.ps}
\vspace{0.5cm}
\caption{The encoding, message corruption in the noisy channel and
decoding can be represented as a Markovian process. The aim is to obtain
a good estimative $\mbox{\boldmath $\widehat {\xi}$}$ for the
original message $\xi$.}
\label{encode}
\end{figure}
In the mapping proposed by Sourlas a message is represented by a
binary vector $\mbox{\boldmath $\xi$}
\in\{\pm 1\}^N$ encoded to a higher dimensional vector
$\mbox{\boldmath $J^0$}\in\{\pm 1\}^M$ defined as $J^{0}_{\langle i_{1},
i_{2} \ldots i_{K}\rangle} = \xi_{i_{1}} \xi_{i_{2}} \ldots \xi_{i_{K}}$,
where $M$ sets of $K$ indices are randomly chosen. A corrupted version
$\mbox{\boldmath $J$}$ of the encoded message $\mbox{\boldmath $J^0$}$ has to
be decoded for retrieving the original message. The decoding process can be
viewed as a statistical Bayesian process \cite{iba98} (see Fig.\ref{encode}).
Decoding focuses on producing an estimate
$\widehat{\mbox{\boldmath $\xi$}}$ to the original message that minimizes a given expected
loss $\langle\langle {\cal L}(\xi,\widehat{\xi})\rangle_{p(J\mid\xi)}\rangle_
{p(\xi)}$ averaged over the indicated probability distributions. The
definition of the loss depends on the particular task; the simple Hamming
distance ${\cal L}(\xi,\widehat{\xi})=\sum_j \xi_j \widehat{\xi}_j$ can be
used for decoding binary messages. An optimal estimator for this particular
loss function is $\widehat{\xi}_j=\mbox{sign}\langle S_j
\rangle_{p(S\mid J)}$ \cite{iba98}, where $\mbox{\boldmath $S$}$
is a $N$ dimensional binary vector representing outcomes of the
decoding process.
Using Bayes' theorem, the posterior
probability can be written as $\mbox{ln }p(\mbox{\boldmath $S$}\mid\mbox{\boldmath $J$} )=\mbox{ln }p(\mbox{\boldmath $J$}\mid \mbox{\boldmath $S$})
+ \mbox{ln }p(\mbox{\boldmath $S$}) + \mbox{const}$.
Sourlas has shown \cite{sourlas94} that for parity check codes this
posterior can be written as a many-body Hamiltonian:
\begin{eqnarray}
\label{eq:Hamiltonian}
\mbox{ln }p(\mbox{\boldmath $S$}\mid\mbox{\boldmath $J$} )&=&-\beta\; {\cal H}(\mbox{\boldmath $S$})\nonumber\\
&=&\beta \sum_{\mu}
{\cal A}_{\mu} \ J_{\mu} \ \prod_{i\in\mu} S_{i} +
\beta{\cal H}_{\mbox{\scriptsize prior}} (\mbox{\boldmath $S$}),
\end{eqnarray}
where $\mu=\left\langle i_{1},\ldots i_{K} \right\rangle$ is a set of indices
and ${\cal A}$ is a tensor with the properties ${\cal A}_\mu\in\{0,1\}$ and
$\sum_{\mu\setminus i}{\cal A}_\mu=C$ $\forall i$, which determines the $M$
components of the codeword $\mbox{\boldmath $J$}^{0}$. The second term ${\cal H}_{\mbox{\scriptsize prior}} (\mbox{\boldmath $S$})$ stands for
the prior knowledge on the actual messages; it can
be chosen as ${\cal H}_{\mbox{\scriptsize prior}}(\mbox{\boldmath $S$})=F\sum_{j=1}^{N} S_j$
to represent the expected bias in the
message bits. For the simple case of a memoryless binary symmetric channel
(BSC), $\mbox{\boldmath $J$}$ is a corrupted version of the
transmitted message $\mbox{\boldmath $J$}^{0}$ where each bit is
independently flipped with probability $p$ during
transmission. The hyper-parameter $\beta$, that reaches
an optimal value at Nishimori's temperature \cite{iba98,rujan93,nishi},
is related to the channel corruption rate.
The decoding procedure translates to finding the thermodynamical spin
averages for the system defined by the Hamiltonian (\ref{eq:Hamiltonian}) at
a certain temperature (Nishimori's temperature for optimal decoding); as the
original message is binary, the retrieved message bits are given by the signs
of the corresponding averages.
In the statistical physics framework the performance of the
error-correcting process can be measured by the overlap between actual
message and estimate for a given scenario characterized by a code rate,
corruption process and information content of the message. To asses
the typical properties we average this overlap over all possible
codes $\cal A$ and noise realizations (possible corrupted vectors
$\mbox{\boldmath $J$}$) given the message $\mbox{\boldmath $\xi$}$ and
then over all possible messages:
\begin{equation}
\label{eq:mag}
m=\frac {1}{N}\left \langle \sum_{i=1}^N {\xi}_i \;\left \langle
\mbox{sign}\langle S_i \rangle \right\rangle_{{\cal A},J|\xi}\right
\rangle_{\xi}
\end{equation}
Here $\mbox{sign}\langle S_i \rangle$ is the sign of the spins thermal
average corresponding to the Bayesian optimal decoding. The average error
per bit is then given by $p_e = (1-m)/2 $. Although this performance measure is not the usual physical magnetization (it can be better described as a measure of misalignment of the decoded message), for brevity,
we will refer to it as {\it magnetization}.
From the statistical physics point of view, the number of checks per bit is
analogous to the spin system connectivity and the number of
bits in each check is analogous to the number of spins per interaction.
Sourlas' code has been studied in the case of extensive connectivity
, where the
number of bonds $C \!\sim\!$ \scriptsize $\left( \begin{array}{c} N-1 \\ K-1
\end{array} \right)$ \normalsize scales with the system size. In
this case it can be mapped onto known problems
in statistical physics such as the SK \cite{SK} ($K\!\!=\!\!2$) and
Random Energy (REM) \cite{Derrida_REM} ($K \!\!\rightarrow\!\!
\infty$) models. It has been shown that the REM saturates
Shannon's bound \cite{sourlas89}. However, it has a rather limited practical
relevance as the choice of extensive connectivity corresponds to a
vanishingly small code rate.
Here we present an analysis of Sourlas' code for the case of finite
connectivity where the code rate is non-vanishing,
detailing and extending our previous brief reports \cite{ks98a,ks98b}.
We show that Shannon's bound can also be attained at finite code rates.
We study the decoding dynamics and discuss the connections between
statistical physics and belief propagation methods.
This paper is organized as follows: in Section II we introduce a naive
mean-field
model that contains all the necessary ingredients to understand the system qualitatively. Section III describes the statistical
physics treatment of Sourlas' code showing that Shannon's bound
can be attained for finite code rates if $K\rightarrow\infty$. The finite $K$ case and the Gaussian noise are also discussed in Section III. The decoding dynamics is analyzed in Section IV. Concluding remarks are given in
Section V. Appendices with detailed calculations are also provided.
\section{Naive Mean Field Theory}
\label{sec:naive}
\subsection {Equilibrium}
\begin{figure}
\hspace*{.4cm}
\epsfxsize=120mm \epsfbox{figure2.eps}
\caption{Code performance measured by the magnetization $m$ as a function
of the noise level $p$ as given by the naive mean-field theory at code rate
$R=1/2$ and $K=2,3,4$ respectively from the bottom. The long-dashed line
indicates PARA-FERRO coexistence. Insets: Maximum initial
deviation $\lambda$ for convergence at a noise level $p=0.1$. Top inset:
$K=3$ and increasing $C$. Bottom inset: Code rate $R=1/2$ and
increasing $K$.}
\label{naive}
\end{figure}
To gain some insight into the code behavior one can start by considering
that the original message is $\xi_j=1$ for all $j$ (so $m=1$ will correspond
to perfect decoding) and use Weiss' mean-field theory as a first (naive)
approximation. The idea is to consider an
effective field given by (for unbiased messages with $F=0$):
\begin {equation}
\label{eq:naiveeff}
h^{\mbox{\scriptsize eff}}_j =\sum_{\{\mu:j\in \mu\}}J_\mu \prod_{i\in\mu\setminus j} S_i
\end{equation}
acting in every site. The first strong approximation here consists in
disregarding
the reaction fields that describe the influence of site $j$ back over the
system.
The local magnetization can then be calculated:
\begin {equation}
\label{eq:naivemag}
m_j=\left\langle\mbox{tanh}\left(\beta h^{\mbox{\scriptsize eff}}_j \right)\right\rangle_{J,S}\simeq \mbox{tanh}\beta \left\langle h^{\mbox{\scriptsize eff}}_j \right\rangle_{J,S},
\end{equation}
where we introduced a further approximation taking averages inside the
function that can be seen as a high temperature approximation. Disregarding correlations among spins and computing
the proper averages one can write:
\begin {equation}
\label{eq:naivemagav}
m=\mbox{tanh} \left( \beta \; C (1-2p)\; m^{K-1}\right),
\end{equation}
where $p$ is the noise level in the channel. An alternative way to derive the above equation is by considering the free-energy:
\begin {equation}
\label{eq:naivefree}
f(m)=-(1-2p)\frac{C}{K}m^{K}-\frac{s(m)}{\beta}.
\end{equation}
The entropic term $s(m)$ is:
\begin {equation}
s(m)=-\frac{1+m}{2}\mbox{ ln}
\left(\frac{1+m}{2}\right)-\frac{1-m}{2}\mbox{ ln}
\left(\frac{1-m}{2}\right).
\end{equation}
Minimizing this free-energy one can obtain Eq.(\ref{eq:naivemagav}) whose
solutions give the possible phases after the decoding process. In Fig.
\ref{naive} we show the maximum magnetization solutions $m$ for
Eq.(\ref{eq:naivemagav}) as a function of the flip rate $p$ at code rate
$R=1/2$ and $K=2,3,4$. For $K=2$ the performance degrades faster with the
noise level than in the $K>2$ case. The dashed line indicates coexistence
between paramagnetic (PARA) $m=0$ and ferromagnetic (FERRO) $m>0$ phases.
\subsection{Decoding Dynamics}
\begin{figure}
\hspace*{3cm}
\epsfxsize=50mm \epsfbox{figure3.ps}
\vspace{0.5cm}
\caption{Graph representing a code.}
\label{node}
\end{figure}
In a naive mean-field framework the decoding process can be
seen as an iterative solution for (\ref{eq:naivemagav}) starting from a
magnetization value that depends on the prior knowledge about the
original message. The fixed points of this dynamics correspond to the minima of the free-energy; a specific minimum is reached depending on the initial condition. In the insets of Fig.\ref{naive} we show, as a measure for the basin of attraction, the maximal deviation between
the initial condition and the original message $\lambda=1-m_0$ that allows
convergence to a FERRO solution. At the bottom inset we show the deviation
$\lambda$ at code rate $R=1/2$, increasing values of $K$ and
noise level $p=0.1$ .
An increasing initial magnetization is needed when $K$ increases,
decoding without prior knowledge is only possible for $K=2$. The top
inset shows $\lambda$ for $K=3$, $p=0.1$; as $C$ increases (code rate decreases), the basin of attraction increases.
One can understand intuitively how the basin of attraction depends on
the connectivities by representing the code in a graph with bit and check nodes and looking at the mean-field
behavior of a single bit node (see Fig.\ref{node}). The corrupted checks
contribute wrong ($-1$ for the ``all ones'' message case) values to the bit nodes ($m<1$ in the mean field). Since check node values correspond to a product of $K-1$ bit values, the probability of updating these nodes to the wrong values increases with $K$, degrading
the overall performance. On the other hand, if $C$ increases for a fixed $K$ the bit nodes gather more information and are less sensitive to the presence of (a limited amount of ) wrong bits .
Although this naive picture indicates some of the qualitative features
of real codes, one certainly cannot rely in its numerical predictions.
In the following sections we will study Sourlas' codes using more
sophisticated techniques that will substantially refine the analysis.
\section{Equilibrium}
\subsection{Replica Theory}
\label{sec:replica}
In the following subsections we will develop the replica symmetric theory for Sourlas' codes and show that, in addition to providing a good description of the equilibrium, it describes the typical decoding dynamics using
belief propagation methods.
The previous naive ``all ones'' messages assumption can be formally translated
to the gauge transformation \cite{frad} $S_{i} \!\!\mapsto\!\!
S_{i} \xi_{i}$ and $J_{\mu}\!\!\mapsto\!\! J_{\mu}\prod_{i\in \mu} \xi_{i} $
that maps any general message to the FERRO configuration defined as
$\xi_i^{*}=1$ $\forall i$. One can then rewrite the Hamiltonian in the form:
\begin{equation}
\label{eq:Hamiltonian_gauge} {\cal H}(\mbox{\boldmath $S$})=-
\sum_{\mu} {\cal A}_{\mu} \ J_{\mu} \ \prod_{i\in\mu} S_{i} -
F \sum_{k}\xi_k S_{k} \ ,
\end{equation}
With this transformation, the bits of the uncorrupted encoded message are
$J^0_i=1$ $\forall i$ and, for a BSC, the corrupted bits are random variables with probability:
\begin{equation}
\label{eq:xi_J_prob_dist}
{\cal P}\left(J_{\mu}\right) = (1\!-\!p)\ \delta \left(J_{\mu} \!-\! 1 \right) + p \ \delta\left(J_{\mu} \!+\! 1 \right),
\end{equation}
where $p$ is the channel flip rate. For deriving typical properties of these codes one has obtain an expression for the free-energy by invoking the replica approach where the free-energy is defined as:
\begin{equation}
\label{eq:freenergy}
f= -\frac{1}{\beta}\lim_{N\rightarrow\infty} \frac{1}{N}
\left.\frac{\partial} {\partial {\mathit n}}\right |_{{\mathit n}=0} \langle {\cal Z}^{ \mathit n}\rangle_{{\cal A},\xi,J},
\end{equation}
where $\langle {\cal Z}^{ \mathit n}\rangle_{{\cal A},\xi,J}$ represents
an analytical continuation in the interval $n\in[0,1]$ of the replicated
partition function defined as:
\begin{equation}
\label{eq:partit}
\langle {\cal Z}^{n}\rangle_{{\cal A},\xi,J} = \mbox{Tr}_{\{S_j^\alpha\}}
\left[\left \langle e^{ \beta F \sum_{\alpha,k}\xi_k S^\alpha_{k}}\right
\rangle_{\xi}\left\langle \exp\left(\beta\sum_{\alpha,\mu}
{\cal A}_{\mu} \ J_{\mu} \ \prod_{i\in\mu} S^\alpha_{i} \right)
\right\rangle_{{\cal A},J} \right].
\end{equation}
The magnetization can be
rewritten in the gauged variables as :
\begin{equation}
\label{eq:mag_gauged}
m= \left \langle\left \langle \mbox{sign}\langle S_i \rangle
\right\rangle_{{\cal A},J|\xi^*}\right \rangle_{\xi},
\end{equation}
where $\xi^*$ denotes the transformation of a message $\xi$ into
the FERRO configuration. The usual magnetization per site can be
easily obtained by calculating
\begin{equation}
\label{eq:fundamental}
\left \langle\left \langle S_i \right \rangle\right\rangle_{{\cal A},J,\xi}=- \left( \frac {\partial f} { \partial(\xi F)} \right).
\end{equation}
From this derivative one can find the distribution of the effective local
fields $h_j$ that can be used to asses the magnetization $m$, since
$\mbox{sign}\left(\langle S_j\rangle\right)=\mbox{sign}(h_j)$ .
To compute the replicated partition function we closely follow
Ref. \cite{wong_a}. We average uniformly over all codes ${\cal A}$ such
that $\sum_{\mu\setminus i}{\cal A}_{\mu}= C$ $\forall i$ to find:
\begin{eqnarray}
\label{eq:partit_2}
\langle {\cal Z}^{n}\rangle_{{\cal A},\xi,J}& =&\exp \left\{ N \;Extr_
{q,\widehat{q}}\left[C-\frac{C}{K}+\frac{C}{K}\left(\sum_{l=0}^{n}{\cal T}_l\sum_{\langle \alpha_1 \ldots \alpha_l\rangle}
q_{\alpha_1 \ldots \alpha_l}^{K} \right)\right.\right.\nonumber\\
& -&\left.\left. C \left(\sum_{l=0}^{n}\sum_{\langle \alpha_1 \ldots
\alpha_l\rangle}q_{\alpha_1 \ldots \alpha_l}\widehat{q}
_{\alpha_1 \ldots \alpha_l}\right) \nonumber \right. \right.\\
&+& \left.\left.\ln \mbox{Tr}_{\{S^{\alpha}\}}\left \langle e^{\beta F\xi\sum
_{\alpha}S^\alpha}\right\rangle_{\xi}\left(\sum_{l=0}^{n}
\sum_{\langle \alpha_1 \ldots \alpha_l\rangle}\widehat{q}
_{\alpha_1 \ldots \alpha_l}S^{\alpha_1}\ldots S^{\alpha_l} \right)^C \right]\right\},
\end{eqnarray}
where ${\cal T}_l=\langle \tanh^l(\beta J) \rangle_J$, as
in \cite{viana}, and $q_0=1$. We give details of this calculation in
the Appendix A.
At the extremum the order parameters acquire expressions similar to those of
Ref. \cite{wong_a}:
\begin{eqnarray}
\label{order-param}
\widehat{q}_{\alpha_1,...,\alpha_l}&=& {\cal T}_l\; q^{K-1}_
{\alpha_1,...,\alpha_l}\nonumber \\
q_{\alpha_1,...,\alpha_l}&=&\left \langle \left (\prod_{i=1}^l
S^{\alpha_i} \right) \left(\sum_{l=0}^{n}\sum_{\langle \alpha_1 \ldots
\alpha_l\rangle}\widehat{q}_{\alpha_1 \ldots \alpha_l}S^
{\alpha_1}\ldots S^{\alpha_l}\right)^{-1}\right\rangle_{\cal X}.
\end{eqnarray}
where
\begin{equation}
{\cal X}=\left \langle e^{\beta F\xi\sum_{\alpha}S^\alpha}\right\rangle_
{\xi}\left(\sum_{l=0}^{n}\sum_{\langle
\alpha_1 \ldots\alpha_l\rangle}\widehat{q}_{\alpha_1 \ldots \alpha_l}
S^{\alpha_1}\ldots S^{\alpha_l} \right)^{C},
\end{equation}
and $\langle...\rangle_{\cal X}=\mbox{Tr}_{\{S^{\alpha}\}}\left[(...)
{\cal X}\right]/\mbox{Tr}_{\{S^{\alpha}\}}\left[(...)\right]$.
The term
$\widehat{p}(\underline {S})=\sum_{l=0}^{n}\sum_{\langle \alpha_1 \ldots\alpha_l\rangle}
\widehat{q}_{\alpha_1 \ldots \alpha_l} S^{\alpha_1}\ldots S^{\alpha_l}$
represents a probability distribution
over the space of replicas and $p_0(\underline{S})=\left \langle e^{\beta F\xi\sum_{\alpha}S^\alpha}\right\rangle_{\xi}$ is a prior distribution over the same space. For reasons that will become clear in
Section \ref{sec:decoding}, $q_{\alpha_1,...,\alpha_l}$ represents one
$l$-th momentum of the
equilibrium distribution of a bit-check edge in a belief network during the
decoding process and $\widehat{q}_{\alpha_1 \ldots \alpha_l}$
represents $l$-th moments of a check-bit edge
equilibrium distribution . The distribution ${\cal X}$ represents the probability of a certain site (bit node) configuration subjected to exactly $C$
interactions and with prior probability given by $p_0$.
\subsection{Replica Symmetric Solution}
\label{sec:symmetric}
The replica symmetric (RS) ansatz can be introduced via the auxiliary
fields $\pi(x)$ and $\widehat{\pi}(y)$ in the following way
(see also \cite{wong_a}):
\begin{eqnarray}
\label{eq:auxfields}
\widehat{q}_{\alpha_1 ... \alpha_l}&=&\int \: dy \; \widehat{\pi}(y)
\tanh^l(\beta y) ,\nonumber\\
q_{\alpha_1 ... \alpha_l}&=&\int \: dx \; \pi(x) \tanh^l(\beta x)
\end{eqnarray}
for $l=1,2,\ldots$.
Plugging it into the replicated partition function (\ref{eq:partit_2}),
performing the limit $n\rightarrow 0$ and using Eq.(\ref{eq:freenergy})
(see Appendix \ref{app:B} for details) one obtains:
\begin{eqnarray}
\label{eq:freesym}
f&=&-\frac{1}{\beta}\: Extr_{\pi,\widehat{\pi}}\left \{\alpha \ln \cosh
\beta \right. \\
&+& \alpha \int \left[
\prod_{l=1}^{K} dx_{l} \ \pi(x_{l}) \right] \left\langle \ln \left[ 1
+ \tanh \beta J \ \prod_{j=1}^{K} \tanh \beta x_{j} \right]
\right\rangle_{J} \nonumber \\
&-& C \int dx \ dy \ \pi(x) \
\widehat{\pi}(y) \ \ln \left[ 1 + \tanh \beta x \ \tanh \beta y
\right] \nonumber\\
&-& C \int dy \ \widehat{\pi}(y) \ \ln \cosh \beta y \nonumber\\
&+&\left. \int \left[ \prod_{l=1}^{C} dy_{l} \ \widehat{\pi}(y_{l})
\right] \left\langle \ln \left[ 2 \cosh \beta \left(\sum_{j=1}^{C}
y_{j} + F \xi \right) \right]\right\rangle_{\xi} \right \}\nonumber,
\end{eqnarray}
where $\alpha=C/K$.
The saddle-point equations, obtained by varying
Eq.(\ref{eq:freesym}) with respect to the probability distributions,
provide a set of relations between $\pi(x)$ and $\widehat{\pi}(y)$
\begin{eqnarray}
\label{eq:saddle_point}
\fl\pi(x) &=& \int \left[ \prod_{l=1}^{C-1} dy_{l} \ \widehat{\pi}(y_{l})
\right] \ \left\langle \delta \left[ x - \sum_{j=1}^{C-1} y_{j} - F \xi
\right]\right\rangle_{\xi} \\
\fl \widehat{\pi}(y) &=& \int \left[
\prod_{l=1}^{K-1} dx_{l} \ \pi(x_{l}) \right] \ \left\langle \delta
\left[ y -
\frac{1}{\beta} \tanh^{-1} \left( \tanh\beta J \ \prod_{j=1}^{K-1}
\tanh \beta x_{j} \right) \right] \right\rangle_{J} \ . \nonumber
\end{eqnarray}
Later we will show that this self-consistent pair of equations can be seen as
a mean-field version for the belief propagation decoding.
Using Eq.(\ref{eq:fundamental}) one finds that the local field distribution is
:
\begin{equation}
\label{eq:local_field}
P(h)=\int \left[ \prod_{l=1}^{C} dy_{l} \
\widehat{\pi}(y_{l}) \right] \ \left\langle \delta \left[ h -
\sum_{j=1}^{C} y_{j} - F \xi \right]\right\rangle_{\xi},
\end{equation}
where $\widehat{\pi}(y)$ is given by the saddle point equations above.
The magnetization (\ref{eq:mag}) can then be calculated using:
\begin{equation}
\label{eq:mag_sym}
m = \int d h \ \mbox{sign} (h) \, P(h).
\end{equation}
The code performance can be assessed by assuming a particular prior distribution for the message bits,
solving the saddle-point equations (\ref{eq:saddle_point}) numerically and
then computing the magnetization.
Instabilities in the solution within the space of symmetric replicas can be
probed looking at second derivatives of the functional whose extremum
defines the free-energy (\ref{eq:freesym}). The simplest necessary
condition for stability is having non-negative second functional derivatives
in relation to $\pi(x)$ (and $\widehat{\pi}(y)$) :
\begin{equation}
\label{eq:stability}
\fl \frac{1}{\beta} \int \left[\prod_{l=1}^{K-2} dx_{l} \
\pi(x_{l}) \right] \left\langle
\ln \left[ 1+ \tanh \beta J \ \tanh^2 \beta x\prod_{j=1}^{K-2} \tanh
\beta x_{j} \right]
\right\rangle_{J} \geq 0,
\end{equation}
for all $x$.
The replica symmetric solution is expected to be unstable for
sufficiently low temperatures (large $\beta$). For high temperatures
we can expand the above expression around small $\beta$ to find the
stability condition:
\begin{equation}
\label{eq:stabhigh}
\langle J\rangle_{J} \langle x \rangle_{\pi}^{K-2}\geq 0 \;
\end{equation}
We expect the average $\langle x \rangle_{\pi}=\int dx\, \pi(x)\, x$
to be zero in PARA phase and positive in FERRO phase,
satisfying the stability condition. This result is still generally inconclusive, but provides some evidence that can be examined numerically. In Section \ref{sec:bound} we will test the stability of our solutions using
condition (\ref{eq:stability}).
In the next sections we restrict our study to the unbiased case ($F=0$),
which is of practical relevance, since it is always possible to compress a biased message to an unbiased one.
\subsection {Case $K\rightarrow\infty$, $C=\alpha K$ }
\label{sec:Kinfty}
For this case one can obtain solutions to the
saddle-point equations for arbitrary temperatures. In the first saddle-point
equation (\ref{eq:saddle_point}) one can write:
\begin{equation}
\label{central_limit}
x=\sum_{l=1}^{C-1} y_l \approx (C-1)\langle y \rangle_{\widehat{\pi}} = (C-1)\int dy \; y\;
\widehat{\pi}(y).
\end{equation}
It means that if $\langle y \rangle_{\widehat{\pi}}=0$ (as it is the in
PARA and spin glass (SG) phases)
then $\pi(x)$ must be concentrated at $x=0$ implying that
${\pi}(x)=\delta(x)$ and $\widehat{\pi}(y)=\delta(y)$ are
the only possible solutions. Moreover, Eq.(\ref{central_limit})
implies that in FERRO phase one can expect
$x\approx{\cal O}(K)$.
Using Eq.(\ref{central_limit}) and the second saddle-point equation
(\ref{eq:saddle_point}) one can find a self-consistent equation for the
mean-field $\langle y \rangle_{\widehat{\pi}}$:
\begin{equation}
\label{mean_field}
\langle y \rangle_{\widehat{\pi}} = \left \langle\frac{1}{\beta}\, \mbox{tanh}^{-1}
\left[\mbox{tanh}(\beta J)\left(\mbox{tanh}(\beta (C-1)\langle y \rangle_{\widehat{\pi}})
\right)^{K-1} \right]\right\rangle_J.
\end{equation}
For a BSC the above average is over distribution
(\ref{eq:xi_J_prob_dist}). Computing the average, using $C=\alpha K$
and rescaling the temperature as $\beta = \tilde{\beta} (\mbox {ln}K)/K $,
in the limit $K\rightarrow\infty$ one obtains:
\begin{equation}
\label{mean_field2}
\langle y \rangle_{\widehat{\pi}} = (1-2p)\left[\mbox{tanh}
(\tilde{\beta}\alpha\langle y\rangle_{\widehat{\pi}}
\mbox{ ln}(K)) \right]^{K},
\end{equation}
where $p$ is the channel flip probability.
The mean-field $\langle y \rangle_{\widehat{\pi}} = 0 $ is always a solution
to this equation (either
PARA or SG); at $\beta_c =\mbox{ln}(K) /( 2\alpha K(1-2p))$
an extra non-trivial FERRO solution emerges with
$\langle y \rangle_{\widehat{\pi}}=1-2p$. As the connection with
the magnetization $m$ is given by Eq. (\ref{eq:local_field}) and
Eq. (\ref{eq:mag_sym}); it is not difficult to see that it implies $m=1$ for
FERRO solution. One remarkable
point is that the temperature were the FERRO
solution emerges is
$\beta_c \sim {\cal O}(\mbox{ln}(K)/K)$; it means that in a simulated
annealing process PARA-FERRO barriers emerge quite early for large $K$
values implying metastability and, consequently, a very slow convergence.
It seems to advocate the use of small
$K$ values in practical applications. This case is analyzed in Section
\ref{sec:finite}. For $\beta>\beta_c$ both PARA and FERRO solutions exist.
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure4.eps}
\caption{Phase diagram in the plane temperature $T$ versus noise level $p$
for $K\rightarrow\infty$ and $C=\alpha K$, with
$\alpha=4$. The dotted line indicates Nishimori's temperature $T_N$
. Full lines represent coexistence. The critical noise level is $p_c$.
The necessary condition for stability in the FERRO phase is satisfied above
the dashed line.}
\label{phase}
\end{figure}
The FERRO free-energy can be obtained from Eq.(\ref{eq:freesym}) using Eq.(\ref{central_limit}), being $f_{\mbox{\scriptsize FERRO}}=-\alpha
(1-2p)$. The corresponding entropy is $s_{\mbox{\scriptsize FERRO}}=0$
indicating a single solution. The PARA free-energy is obtained by
plugging $\pi(x)=\delta(x)$ and $\widehat{\pi}(y)=\delta(y)$ into Eq. (\ref{eq:freesym}):
\begin{eqnarray}
\label{eq:freepara}
f_{\mbox{\scriptsize PARA}}&=& -\frac{1}{\beta}(\alpha \mbox{ ln}(\mbox{cosh }\beta) + \mbox{ln }2)
,\\
s_{\mbox{\scriptsize PARA}}&=&\alpha(\mbox{ln}(\mbox{cosh }\beta) -\beta\mbox{ tanh }\beta) +
\mbox{ln }2.
\end{eqnarray}
PARA solutions are unphysical for
$\alpha > (\mbox {ln } 2)/(\beta \mbox { tanh }\beta - \mbox{ln ch }\beta)$,
since the corresponding entropy is negative.
To complete the phase diagram picture we have to assess the spin-glass free-energy and
entropy. We have seen in the beginning of this section that replica symmetric SG and PARA solutions consist of the same field distributions for $K\rightarrow\infty$,
implying unphysical behavior. In order to produce a
solution with non-negative entropy one has to break the replica symmetry.
We use here a pragmatic way to build this solution, using the
simplest one-step replica symmetry breaking known as {\it frozen spins}.
It was observed in Ref. \cite{gross} that for the REM a one-step
symmetry breaking scheme gives the exact solution. In this scheme the $n$
replicas' space is divided to groups of $m$ identical solutions.
It was shown that an abrupt transition in the order parameter from a unique
solution (Edwards-Anderson parameter $q=1$, SG phase) to a completely uncorrelated set of solutions ($q=0$, PARA phase) occurs.
This transition takes place at a critical temperature
$\beta_g$ that can be found by solving the appropriate saddle-point equations;
this temperature is given by the root of the replica symmetric entropy ($s_{\scriptsize RS}=0$) meaning that the RS-RSB transition occurs at the same point as the PARA-SG in this model. The symmetry breaking parameter was found to be
$m_g=\beta_g/\beta$, indicating that this kind of solution
is physical only for $\beta>\beta_g$, since
$m_g \leq 1$ \cite{parisi},
indicating a PARA-SG phase transition. The free-energy can be computed by plugging the order parameters in the effective
Hamiltonian, obtained after averaging over the disorder and taking the proper limits. It shows no dependence on the
temperature, since for $\beta>\beta_g$ the system is completely
frozen in a single configuration.
For the Sourlas' code, in the regime we are interested in, SG solutions
to the saddle-point equations are given by $\pi(x)=\delta(x)$ and
$\widehat{\pi}(y)=\delta(y)$. The RSB-SG free-energy that guaranties
continuity in the SG-PARA transition is identical to $f_{\mbox{\scriptsize PARA}}$,
since the SG and PARA solutions have exactly the same structure, to say:
\begin{equation}
\label{eq:free_rsb_sg}
f_{\mbox{\scriptsize RSB-SG}}=-\frac{1}{\beta_g}\;(\alpha \;\mbox{ln}\,(\mbox{cosh }\,\beta_g) +
\mbox{ln }\,2),
\end{equation}
where $\beta_g$ is a solution for $s_{\mbox{\scriptsize RS-SG}}= \alpha\;(\mbox{ln}\,(\mbox{cosh }\,\beta) -\beta\;\mbox{tanh }\,\beta) +\mbox{ln }\,2=0$.
In Fig.\ref{phase} we show the phase diagram for a given code rate $R$ in
the temperature $T$ versus noise level $p$ plane.
\subsection{Shannon's Limit}
\label{sec:bound}
Shannon's analysis shows that up to a critical code rate $R_c$, which equals
the channel capacity, it is possible to recover information with arbitrarily
small error probability for a given noise level. For the BSC :
\begin{equation}
\label{eq:shannon}
R_c=\frac{1}{\alpha_c}=1+p\mbox{ log}_2 \;p + (1-p)\mbox{ log}_2\; (1-p).
\end{equation}
Sourlas' code, in the case where
$K\rightarrow\infty$ and $C \sim {\cal O}(N^K)$ can be mapped onto the REM
and has been shown to be capable of saturating Shannon's bound in the limit
$R\rightarrow 0$ \cite{sourlas89}. In this section we extend the analysis
to show that Shannon's bound can be attained by Sourlas'
code at zero temperature also for
$K\rightarrow\infty$ limit but with connectivity $C=\alpha K$.
In this limit the model is analogous to the diluted REM analyzed by Saakian
in \cite {saakian}.
The errorless phase is manifested in a FERRO phase with perfect
alignment ($m=1$) (condition that is only possible for infinite $K$) up
to a certain critical noise level; a further noise level increase produces
frustration leading to a SG phase where
the misalignment is maximal ($m=0$). The FERRO-SG transition is
analogous to the transition from errorless decoding to decoding with
errors described by Shannon. A PARA phase is
also present when the transmitted information is insufficient to
recover the original message ($R>1$).
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure5.eps}
\caption{Histogram representing the mean-field distribution
$\widehat{\pi}(y)$ obtained by Monte-carlo integration at low
temperature ($\beta=10$, $K=3$,$C=6$ and $p=0.1$). Dotted lines represent
solutions obtained by iterating self-consistent equations both with five peak and three peak ans\"atze. Inset: detailed view of the weak regular part arising in the Monte-carlo integration. }
\label{fields}
\end{figure}
At zero temperature saddle-point equations (\ref{eq:saddle_point}) can
be rewritten as:
\begin{eqnarray}
\label{eq:sp_infty}
\fl \pi(x) &=& \int \left[ \prod_{l=1}^{C-1} dy_{l} \ \widehat{\pi}(y_{l})\right]
\ \delta \left[ x - \sum_{j=1}^{C-1} y_{j} \right] \\
\fl \widehat{\pi}(y) &=& \int \left[
\prod_{l=1}^{K-1} dx_{l} \ \pi(x_{l}) \right] \ \left\langle \delta
\left[ y - \mbox{sign}(J \prod_{l=1}^{K-1} x_{l}) \mbox{min}(\mid J\mid,
... , \mid x_{K-1} \mid)\right] \right\rangle_{J} \ ,
\nonumber
\end{eqnarray}
The solutions for these saddle-point equations may, in general, result
in probability distributions with singular and regular parts. As a
first approximation we choose the simplest self-consistent family
of solutions which are, since $J=\pm 1$, given by:
\begin{eqnarray}
\widehat{\pi}(y)&=&p_+\delta(y-1)+p_0\delta(y)+p_-\delta(y+1)\\
\pi(x)&=&\sum_{l=1-C}^{C-1} T_{[p_{\pm},p_0;C-1]}(l) \,\delta(x-l),
\end{eqnarray}
with
\begin{equation}
T_{ \left[ p_{+}, p_{0}, p_{-}; C-1 \right]} (l) = \sum^{\prime}_{ \{k,h,m\}} \frac{(C-1)!}{k! \ h! \ m!} \ p_{+}^{k} \
p_{0}^{h} \ p_{-}^{m},
\end{equation}
where the prime indicates that $k,h,m$ are such that $k-h=l; \ k+h+m=C-1$.
Evidence for this simple ansatz comes from Monte-carlo integration of
Equation (\ref{eq:saddle_point}) at very low temperatures, that shows solutions comprising three dominant peaks and a relatively weak regular part.
Inside FERRO and PARA phases a more complex singular solution comprising
five peaks $\widehat{\pi}(y)=p_{+2}\delta(y-1)+p_{+}\delta(y-0.5)+p_0\delta(y)+p_-\delta(y+0.5)+p_{-2}\delta(y+1)$ collapses back to the simpler three peak solution. In Fig.\ref{fields} we show a typical result of a Monte-carlo integration for the field $\widehat{\pi}(y)$. The two peak that emerge by using either the three peak ansatz or the five peak ansatz are shown as dotted lines. In the inset we show the weak regular part of the Monte-carlo solution.
Plugging
the above ansatz in the saddle-point equations one can write a closed set
of equations in $p_{\pm}$ and $p_{0}$ that can be solved numerically
(see appendix D for details).
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure6.eps}
\caption{Phase diagram in the plane code rate $R$ versus noise level $p$
for $K\rightarrow\infty$ and $C=\alpha K$ at zero
temperature. The FERRO-SG coexistence line corresponds to the Shannon's
bound.}
\label{bound}
\end{figure}
The three peak solution can be of three types: FERRO
($p_+>p_-$), PARA ($p_0=1$) or SG ($p_-=p_+$).
Computing free-energies and entropies enables one to construct the phase
diagram. At zero temperature the PARA free-energy is $f_{\mbox{\scriptsize PARA}}=-\alpha$
and the entropy is $s_{\mbox{\scriptsize PARA}}=(1-\alpha)\mbox{ ln }2$, this phase is physical
only for $\alpha<1$, what is expected since it corresponds exactly to the
regime where the transmitted information is not sufficient to recover
the actual message ($R>1$).
The FERRO free-energy does not depend on the temperature, having
the form $f_{\mbox{\scriptsize FERRO}}=-\alpha(1-2p)$ with entropy $s_{\mbox {\scriptsize FERRO}}=0$. One can
find the
FERRO-SG coexistence line that corresponds to the maximum performance of a
Sourlas' code by equating
Eq.(\ref{eq:free_rsb_sg}) and $f_{\mbox{\scriptsize FERRO}}$.
Observing that $\beta_g=\beta_N(p_c)$
(as seen in Fig.\ref{phase} ) we found that this transition
coincides with Shannon's bound
Eq.(\ref{eq:shannon}). It is interesting to note that in the large $K$ regime
both RS-FERRO and RSB-SG free-energies (for $T<T_g$)
do not depend on the temperature, it means that Shannon's bound
is valid also for finite temperatures up to $T_g$.
In Fig.\ref{bound} we give the complete zero temperature phase diagram.
The stability of replica
symmetric FERRO and PARA solutions used to obtain Shannon's bound can
be checked using Eq.(\ref{eq:stability}) at zero temperature:
\begin{equation}
\label{eq:stability0temp}
\fl \lim_{\beta \rightarrow \infty}
\frac{1}{\beta} \int \left[\prod_{l=1}^{K-2} dx_{l} \
\pi(x_{l}) \right] \left\langle
\ln \left[ 1+ \tanh \beta J \ \tanh^2 \beta x\prod_{j=1}^{K-2} \tanh
\beta x_{j} \right]
\right\rangle_{J} \ge 0,
\end{equation}
for all $x$.
For PARA solutions the above integral vanishes, trivially satisfying the
condition, while for FERRO solution in the $K$ large regime,
$x_l\approx{\cal O}(K)$ and the integral becomes
\begin{equation}
-2p \left [ \left (1- \Theta \left( x+1 \right ) \right )
+ |x| \left (\Theta \left( x+1 \right ) -\Theta \left( x- 1 \right ) \right )
+ \Theta \left( x- 1 \right ) \right ],
\end{equation}
where $\Theta(x)=1$ for $x\geq 0$ and $0$ otherwise,
indicating instability for $p>0$.
For the noiseless
case $p=0$ the stability condition is satisfied. The instability of
FERRO phase opens the possibility that Sourlas' code does not saturate
Shannon's bound, since a correction to the
FERRO solution could change FERRO-SG transition line. However, it was
shown in Section \ref{sec:symmetric} that this instability vanishes for large
temperatures, what supports, to some extent, the FERRO-SG line obtained and the saturation
of Shannon's bound in some region, as long as the temperature is lower than Nishimori's temperature.
For finite temperatures the stability condition for FERRO solution
can be rewritten as:
\begin{equation}
\left(1+\mbox{tanh}(\beta)\mbox{tanh}^2(\beta x)\right)^{(1-p)}
\left(1-\mbox{tanh}(\beta)\mbox{tanh}^2(\beta x)\right)^p \ge 1 \; \forall x.
\end{equation}
For $p=0$ the condition is clearly satisfied. For finite $p$
a critical temperature above which the stability condition is fulfilled can be found numerically.
In Fig.\ref{phase} we show this critical temperature in the phase diagram;
one can see that there is a considerable region in which our result that
Sourlas' code can saturate Shannon's bound is supported. Conclusive evidence to
that will be given by simulations presented in Section \ref{sec:decoding}.
\subsection{Finite K Case}
\label{sec:finite}
Although Shannon's bound only can be attained in the limit
$K\rightarrow\infty$, it was shown in the Section \ref{sec:Kinfty}
that there are some possible drawbacks, mainly in the decoding of
messages encoded by large $K$ codes, due to large
barriers which are expected to occur between PARA and FERRO
states. In this section we consider the finite $K$ case,
for which we can solve the RS saddle-point equations (\ref{eq:saddle_point})
for arbitrary
temperatures using Monte-carlo integration. We can also obtain solutions
for the zero temperature case using the simple iterative method described in
Section \ref{sec:bound}.
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure7.eps}
\caption{Top: zero temperature magnetization $m$ as a function of the noise level $p$ for various $K$ values at code rate $R=1/2$, as
obtained by the iterative method
. Notice that the RS theory predicts a transition
of second order for $K=2$ and first order for $K>2$. Bottom:
RS-FERRO free-energies (white circles for $K=2$ and from the
left: $K\,=\,3,4,5$ and $6$) and RSB-SG free-energy (dotted line) as
functions of the noise level $p$.
The arrow indicates the region where the RSB-SG phase starts to dominate.
Inset: a detailed view of the RS-RSB transition region.}
\label{kfinite}
\end{figure}
At the top of Fig.\ref{kfinite} we show the zero temperature
magnetization $m$ as a function of the noise level $p$ at
code rate $R=1/2$.
These curves were obtained by using the
three peak ansatz of the Section \ref{sec:bound}.
It can be seen that the transition is of second order for $K=2$ and first
order for $K>3$ similarly to extensively connected models. The transition
as described by the RS solution tends to $p=0.5$ as $K$ grows. Note that
this does not correspond to perfect retrieval since the RSB spin glass
phase dominates for $p>p_c$ (see bottom of Fig.\ref{kfinite}). In the bottom
figure we plot RS free-energies and RSB frozen
spins free-energy, from which we determine the
critical probability $p_c$ where the transition occurs (pointed by an arrow).
After the transition, free-energies for $K=3,4,5$ and $6$ acquire values
that are lower than the SG free-energy; nevertheless, the entropy
is negative and these free-energies are therefore unphysical. It is
remarkable that this critical value does not change significantly
for finite $K$ in comparison to infinite $K$.
Observe that Shannon's bound
cannot be attained for finite $K$, since $m=1$ exactly only if
$K\rightarrow\infty$.
The $K=2$ model with extensive connectivity (SK) is known to be
somewhat special, a full Parisi solution
is needed to recover the concavity of the free-energy and
the Parisi order function has a continuous behavior \cite{mezard}.
No stable solution
is known for the intensively connected model (Viana-Bray model).
In order to check the theoretical result obtained one relies on
simulations of the decoding process
at low temperatures. In Section VIII we show that the simulations are in good
agreement with the theoretical results.
\subsection {Gaussian Noise}
\label{sec:gauss}
Using the replica symmetric free-energy (\ref{eq:freesym}) and the
frozen spins
RSB free-energy (\ref{eq:free_rsb_sg}) one can easily extend the analysis
to other noise types. The general PARA free-energy and entropy can be written:
\begin{eqnarray}
\label{eq:fparagen}
f_{\mbox{\scriptsize PARA}}&=&-\frac{1}{\beta}\;\left(\alpha\;\langle \mbox{ln}\,(\mbox{ch }\beta J)\rangle_J +\mbox {ln }\,2\right) \nonumber \\
s_{\mbox{\scriptsize PARA}}&=&\alpha\;\left(\langle \mbox{ln}\,(\mbox{ch }\beta J)\rangle_J - \beta \langle J\;\mbox{tanh}\,(\beta J)\rangle_J\right)+\mbox {ln }2.
\end{eqnarray}
The SG-RSB free-energy is given by :
\begin{equation}
\label{eq:fsggen}
f_{\mbox{\scriptsize SG-RSB}}=-\frac{1}{\beta_g}\;\left(\alpha\;\langle \mbox{ln }(\mbox{ch }\beta_g J)\rangle_J +\mbox {ln }2 \right),
\end{equation}
with $\beta_g$ defined as the solution of
\begin{equation}
\label{eq:entropy}
\alpha\;\left(\langle \mbox{ln }(\mbox{ch }\beta_g J)\rangle_J - \beta_g
\langle J\;\mbox{tanh }(\beta_g J)\rangle_J\right)+\mbox {ln }2 =0 .
\end{equation}
The FERRO free-energy is in general given by
$f_{\mbox{\scriptsize FERRO}}=-\alpha\;\langle J\rangle_J=-\alpha\;\langle J \;
\mbox{tanh }(\beta_N J)\rangle_J$ (see Appendix \ref{app:nishifree}).
The maximum performance of the code is defined by the critical line :
\begin{equation}
\label{eq:line}
\alpha\left(\langle \mbox{ln}(\mbox{ch }\beta_g J)\rangle_J - \beta_g
\langle J\;\mbox{tanh}(\beta_N J)\rangle_J\right)+\mbox {ln }2 =0,
\end{equation}
obtained by equating free-energies in PARA and FERRO phases.
Comparing this expression with entropy (\ref{eq:entropy}) it can be
seen that $\beta_g=\beta_N$ at the critical line; the
same behavior observed in the BSC case. From Eq.(\ref{eq:line}) one can
write:
\begin{equation}
\label{eq:capacity}
R_c=\beta_N^2 \frac{\partial}{\partial \beta}
\left[\frac{1}{\beta}\langle \mbox{log}_2
\mbox{ cosh}(\beta J)\rangle_J\right]_{\beta=\beta_N},
\end{equation}
that can be used to compute the performance of the code for arbitrary
symmetric noise.
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure8.eps}
\vspace{1cm}
\caption{Critical code rate $R_c$ and channel capacity for a binary Gaussian
channel as a function of the signal to noise rate $S/N$ (solid line).
Sourlas' code saturates Shannon's bound. Channel capacity of the
unconstrained Gaussian channel (dashed line).}
\label{gaussian}
\end{figure}
Supposing that the encoded bits can acquire totally unconstrained values
Shannon's bound for Gaussian noise is given by
$R_c=\frac{1}{2}\mbox{ log}_2(1 +S/N)$, where $S/N$ is the signal to
noise ratio, defined as the ratio of source energy per bit
(squared amplitude) over the spectral density of the noise (variance).
If one constrains the encoded bits to binary values $\{\pm 1\}$ the capacity of a Gaussian channel
is:
\begin{equation}
\label{eq:capacity_gauss}
R_c=\int dJ\;P(J\mid 1) \mbox{ log}_2 P(J\mid 1) - \int dJ \; P(J)
\mbox{ log}_2 P(J),
\end{equation}
where $P(J\mid J^0)=\frac{1}{\sqrt{2\pi \sigma^2}}\mbox{ exp}(-\frac{(J-J^0)^2}{2\sigma^2})$.
In Fig.\ref{gaussian} we show the performance of Sourlas' code
in a Gaussian channel together with the capacities of the
unconstrained and binary Gaussian channels. We show that
$K\rightarrow\infty$, $C=\alpha K$ Sourlas' code
saturates Shannon's bound for the binary Gaussian channel as well.
The significantly lower performance in the unconstrained
Gaussian channel can be trivially explained by the binary coding
scheme while signal and noise are allowed to acquire real values.
\section {Decoding Dynamics}
\label{sec:decoding}
\subsection{Belief Propagation}
\label{sec:belief}
The decoding process of an error-correcting code relies on
computing averages over the marginal posterior probability
$P(S_j\mid \mbox{\boldmath $J$})$ for each one
of the $N$ message bits $S_j$ given the corrupted encoded bits
$J_\mu$ (checks), where $\mu=\langle i_1 \ldots i_K \rangle$ is one of
the $M$ sets chosen by the tensor ${\cal A}_\mu$.
The probabilistic dependencies
existing in the code can be represented as a bipartite graph known as a
{\it belief network} where nodes in one layer correspond to the $M$
checks $J_\mu$ while nodes in the other to the $N$ bits $S_j$.
Each check is connected
to exactly $K$ bits and each bit is connected exactly to $C$ checks
(see Fig.\ref{belief}a).
Pearl \cite{pearl} proposed an iterative algorithm for computation of
marginal probabilities in belief networks. These algorithms operate by
updating beliefs (conditional probabilities) locally and propagating them.
Generally the convergence of these iterations
depends on the absence of loops in the graph. As can be seen
in Fig.\ref{belief}a, networks that define error-correcting
codes may include loops and convergence problems may occur.
Recently it was shown that in some cases Pearl's algorithm works
even in the presence of loops \cite{weiss}.
\begin{figure}
\hspace*{4cm}
\epsfxsize=80mm \epsfbox{figure9.ps}
\vspace{1cm}
\caption{(a) Belief network representing an error-correcting code.
Each bit $S_j$ (white circles) is linked to exactly $C$ checks and
each check (black circles) $J_{\mu}$ is linked to exactly $K$ bits.
(b) Graphical representation of the field $r_{\mu j}$. The grey box
represents the mean field contribution $\prod_{l\in {\cal L}(\mu)\setminus j}q_{\mu l}$
of the other bits on the check $J_\mu$. (c) Representation of one of the
fields $q_{\mu l}$. }
\label{belief}
\end{figure}
The particular use of belief networks as decoding
algorithms for error-correcting codes based
on sparse matrices was discussed by MacKay in \cite{mackay95b}.
In this work a loop-free approximation for the graph in Fig.\ref{belief}a
was proposed (see \cite{pearl} for a general discussion
on such approximations). In fact, it was shown in \cite{urbanke} that the probability of finite length loops in these graphs vanishes with the system size.
In this framework the
network is decomposed in a way to avoid loops and the conditional probabilities
$q^{(S)}_{\mu j}$ and $r^{(S)}_{\mu j}$ are computed.
The set of bits in a check $\mu$ is defined as ${\cal L}(\mu)$ and the set of checks over the bit $j$ as ${\cal M}(j)$. The probability that $S_j=S$
given information on all checks other than $\mu$ is denoted
$q^{(S)}_{\mu j} =P(S_j=S\mid \{J_{\nu}:\nu \in {\cal M}(j)\setminus\mu\})$
and $r^{(S)}_{\mu j} = \mbox{Tr}_{\{S_l:l\in{\cal L}(\mu)\setminus j\}}
P(J_{\mu}\mid S_j=S, \{S_l:l \in {\cal L}(\mu)\setminus j \})
\prod _{l\in {\cal L}(\mu)\setminus l} q^{(S_l)}_{\mu l}$ is the probability of the check $J_{\mu}$ if the bit $j$ is fixed to $S_j=S$ and the other bits
involved are supposed to have distributions given by $q^{(S_i)}_{\mu i}$
. In Fig.\ref{belief}b one can see a graphical representation of
$r^{(S)}_{\mu j}$ that can be interpreted as the influence of the bit $S_j$ and the mean-field $\prod _{l\in {\cal L}(\mu)\setminus l} q^{(S_l)}_{\mu l}$
(representing bits in ${\cal L}(\mu)$ over than $l$) over the check $J_\mu$. In the Fig.\ref{belief}c
we see that each field $q^{(S)}_{\mu l}$ represents the influence of the
checks in ${\cal M}(l)$, excluding $\mu$, over each bit $S_l$, this
setup excludes the loops that may exist in the actual network.
Employing Bayes theorem, $q^{(S)}_{\mu j}$ can be rewritten as:
\begin{equation}
q^{(S)}_{\mu j} = a_{\mu j}\;P(\{J_{\nu}:\nu \in {\cal M}(j)\setminus\mu\}\mid S_j)\;p^{(S)}_{j},
\end{equation}
where $a_{\mu j}$ is a normalization constant such that
$q^{(+1)}_{\mu j}+q^{(-1)}_{\mu j}=1$ and $p^{(S)}_{j}$ is the prior
probability over the bit $j$. The distribution $P(\{J_{\nu}:\nu \in {\cal M}(j)\setminus\mu\}\mid S_j)$ can be replaced by a mean-field approximation by
factorizing dependencies using fields $r^{(S)}_{\mu j}$:
\begin{eqnarray}
\label{eq:belief}
q^{(S)}_{\mu j}&=&a_{\mu j}p^{(S)}_{j}\prod_{\nu\in{\cal M}(j)\setminus\mu}
r^{(S)}_{\nu j} \nonumber \\
r^{(S)}_{\mu j} &=& \mbox{Tr}_{\{S_l:l\in{\cal L}(\mu)\setminus j\}}
P(J_{\mu}\mid S_j=S, \{S_i:i \in {\cal L}(\mu)\setminus j \})
\prod _{i\in {\cal L}(\mu)\setminus j} q^{(S_i)}_{\mu i}.
\end{eqnarray}
A message estimate
$\widehat\xi_j =\mbox{sign}\left(\langle S_j \rangle_{q^{(S)}_j}\right)$
can be obtained by solving the above equations and
computing the pseudo-posterior:
\begin{equation}
q^{(S)}_{j}=a_{j}p^{(S)}_j\prod_{\nu\in{\cal M}(j)}r^{(S)}_{\nu j},
\label{eq:pseudo}
\end{equation}
where $a_{j}$ is a normalization constant.
By taking advantage of the normalization conditions for the distributions
$q^{(+1)}_{\mu j}+q^{(-1)}_{\mu j}=1$ and
$r^{(+1)}_{\mu j}+r^{(-1)}_{\mu j}=1$
one can change variables and reduce the number of equations (\ref{eq:belief}) to the couple
$\delta q_{\mu j} = q^{(+1)}_{\mu j}-q^{(-1)}_{\mu j}$ and
$\delta r_{\mu j} = r^{(+1)}_{\mu j}-r^{(-1)}_{\mu j}$.
Solving these equations, one can find back $r^{(S)}_{\mu j}=\frac{1}{2}
\left(1\;+\;\delta r_{\mu j}S_j \right)$ and the
pseudo-posterior can be calculated to obtain the estimate.
\subsection{Connection with Statistical Physics}
\label{sec:connection}
The belief propagation algorithm was shown in \cite{mackay95b} to outperform
other methods such as simulated annealing. In \cite{ks98a} it was proposed
that this framework can be reinterpreted using statistical physics.
The main ideas behind the approximations contained in (\ref {eq:belief}) are somewhat similar to the Bethe \cite{bethe} approximation to diluted
two-body spin glasses. Actually,
for systems involving two-body interactions it is known that the Bethe
approximation is equivalent to solving exactly a model defined on a
Cayley tree and that this is a good approximation
for finitely connected systems in the thermodynamical limit \cite{wong_d}.
In fact, loops in the connections become rare as the system size grows
and can be neglected without introducing significant errors. The belief propagation can be seen as a Bethe-like approximation for multiple bodies interaction systems.
The mean-field approximations used here are also quite similar
to the TAP approach \cite{tap}. The fields $q^{(S)}_{\mu j}$ correspond
to the mean influence of other sites other the site $j$ and
the fields $r^{(S)}_{\nu j}$ represent the influence of $j$ back over the system (reaction fields).
The analogy can be exposed by observing that the likelihood
$p(J_\mu\mid \mbox{\boldmath $S$})$
is proportional to the Boltzmann weight:
\begin{equation}
w_B(J_\mu \mid \{S_j:j \in {\cal L}(\mu) \}) = \mbox{exp}\left(-\beta J_\mu \; \prod_{i\in\mu} S_{i}\right).
\end{equation}
That can be also written in the more convenient form:
\begin{equation}
\label{eq:likelihood}
w_B(J_\mu \mid \{S_j :j \in {\cal L}(\mu) \}) = \frac{1}{2} \mbox{cosh}(\beta J_\mu)\left( 1\; +
\;\mbox{tanh}(\beta J_{\mu})\; \prod_{j\in{\cal L}(\mu)}S_j \right).
\end{equation}
The variable $r_{\mu j}^{(S_j)}$ can then be seen as proportional to the effective Boltzmann weight obtained by fixing the bit $S_j$:
\begin{equation}
\label{eq:effective}
w_{\mbox{\scriptsize eff}}(J_\mu \mid S_j)
= \mbox{Tr}_{\{S_l\; :\;l \in {\cal L}(\mu)\setminus j \}}\;
w_B(J_\mu \mid \{S_l\; :\;l \in {\cal L}(\mu) \})\prod _{l\in {\cal L}(\mu)\setminus j}
q^{(S_l)}_{\mu l}.
\end{equation}
Plugging Eq.(\ref{eq:likelihood}) for the likelihood in equations
(\ref{eq:belief}), using the fact that the prior probability is given by $p_j^{(S)}=\frac{1}{2}\left(1+\mbox{tanh}(\beta S F)\right)$ and computing $\delta q_{\mu j}$ and $\delta r_{\mu j}$:
\begin{eqnarray}
\label{eq:tap}
\delta r_{\mu j}&=&\mbox{tanh}(\beta J_\mu) \prod_{l\in{\cal L}(\mu)\setminus j} \delta q_{\mu l}\nonumber \\
\delta q_{\mu j}&=&\mbox{tanh}\left(\sum_{\nu\in{\cal M}(l)\setminus \mu}
\mbox{tanh}^{-1}( \delta r_{\nu j}) +\beta F \right).
\end{eqnarray}
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure10.eps}
\caption{Magnetization as a function of the flip probability $p$ for
decoding using TAP equations for $K=2$. From the bottom:
Monte-carlo solution of the RS saddle-point equations
for unbiased messages ($f_s=0.5$) at $T=0.26$ (line) and
$10$ independent runs of TAP decoding for each flip probability (plus signs),
$T=0.26$ and biased messages ($f_s=0.1$) at
Nishimori's temperature $T_N$.}
\label{TAP1}
\end{figure}
The pseudo-posterior can then be calculated:
\begin{equation}
\label{eq:pseudoposterior}
\delta q_{j}=\mbox{tanh}\left(\sum_{\nu\in{\cal M}(l)}
\mbox{tanh}^{-1}( \delta r_{\nu j}) +\beta F \right),
\end{equation}
providing Bayes' optimal decoding
$\widehat{\xi}_j=\mbox{sign}(\delta q_{j})$. It is
important at this point to support the mean-field
assumptions used here by methods of statistical
physics \cite{ks98a}. The factorizability of the probability
distributions can be explained by weak correlations between connections
(checks) and by the cluster property:
\begin{equation}
\lim_{N\rightarrow \infty} \frac{1}{N^2}\sum_{i\neq j}
\left(\langle S_i S_j\rangle_{p(S\mid J)} -
\langle S_i \rangle_{p(S\mid J)}
\langle S_j \rangle_{p(S\mid J)}\right)^2 \rightarrow 0
\end{equation}
that bits $S_j$ obey within a pure state \cite{mezard}.
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure11.eps}
\caption{Magnetization as a function of the flip probability $p$ for
decoding using TAP equations for $K=5$.
The dotted line is the replica symmetric saddle-point equations Monte-carlo
integration for unbiased messages ($f_s=0.5$) at the Nishimori's temperature $T_N$. The
bottom error bars correspond to $10$ simulations using the TAP decoding.
The decoding performs badly on average in this scenario. The upper curves
are for biased messages ($f_s=0.1$) at the Nishimori's temperature $T_N$.
The simulations agree with results obtained using the replica symmetric ansatz
and Monte-carlo integration.}
\label{TAP2}
\end{figure}
One can push the above connections even further. Eqs.(\ref{eq:tap}), of course, depend on the particular received message $\mbox{\boldmath $J$}$. In order to make the analysis message independent, one can use
a gauge transformation $\delta r_{\mu j} \mapsto \xi_j \delta r_{\mu j}$ and
$\delta q_{\mu j} \mapsto \xi_j \delta q_{\mu j}$ to write:
\begin{eqnarray}
\label{eq:gaugetap}
\delta r_{\mu j}&=&\mbox{tanh}(\beta J) \prod_{l\in{\cal L}(\mu)\setminus j} \delta q_{\mu l}\nonumber \\
\delta q_{\mu j}&=&\mbox{tanh}\left(\sum_{\nu\in{\cal M}(l)\setminus \mu}
\mbox{tanh}^{-1}( \delta r_{\nu j}) +\beta \xi_j F \right).
\end{eqnarray}
In this form a success in the decoding process correspond to $\delta r_{\mu j}>0$ and $\delta q_{\mu j}=1$ for all $\mu$ and $j$. For a large number of iterations, one can expect the ensemble of belief networks to converge to an equilibrium distribution where $\delta r$ and $\delta q$ are random variables sampled from distributions $\widehat{\rho}(y)$ and $\rho(x)$ respectively.
By transforming these variables as $\delta r=\mbox{tanh}(\beta y)$ and
$\delta q=\mbox{tanh}(\beta x)$ and considering the actual message and noise as quenched disorder, Eqs.(\ref{eq:gaugetap}) can be rewritten as:
\begin{eqnarray}
\label{eq:newgaugetap}
y&=&\frac{1}{\beta}\left\langle \mbox{tanh}^{-1}\left(\mbox{tanh}(\beta J) \prod_{j=1}^{K-1} \mbox{tanh}(\beta x_j)\right)\right\rangle_J\nonumber \\
x&=&\left\langle \sum_{j=1}^{C-1} y_j + \xi F\right\rangle_\xi.
\end{eqnarray}
The above relations lead to a dynamics on the distributions $\widehat{\rho}(y)$ and $\rho(x)$, that is exactly the same obtained when solving iteratively
RS saddle-point equations (\ref{eq:saddle_point}). The probability distributions $\widehat{\rho}(y)$ and $\rho(x)$ can be ,therefore, identified with $\widehat{\pi}(y)$ and $\pi(x)$ respectively and the RS solutions correspond to decoding a generic message using belief propagation averaged over an ensemble of different codes, noise and signals.
Eqs.(\ref{eq:tap}) are now used to show the agreement between the simulated
decoding and analytical calculations.
For each run, a fixed
code is used to generate 20000 bit codewords from 10000 bit messages,
corrupted versions of the codewords are then decoded using (\ref{eq:tap}).
Numerical solutions for 10 individual
runs are presented in Figs.\ref{TAP1} and \ref{TAP2},
initial conditions are chosen as $\delta r_{\mu l}=0$
and $\delta q_{\mu l}=\mbox{tanh}(\beta F)$ reflecting prior
beliefs. In Fig.\ref{TAP1}
we show results for $K=2$ and $C=4$
in the unbiased case, at code rate $R=1/2$
(prior probability $p_j^{(1)}=f=0.5$)
at a low temperature $T=0.26$ (we avoided $T=0$ due to numerical
difficulties).
Solving saddle-point equations (\ref{eq:saddle_point}) numerically using
Monte-carlo integration methods we obtain solutions with good agreement to
simulated decoding. In the same figure we show the performance for the case of
biased messages ($p_j^{(1)}=f_s=0.1$), at code rate $R=1/4$. Also here the
agreement with Monte-carlo integrations is rather convincing. The third
curve in Fig.\ref{TAP1} shows the performance
for biased messages at Nishimori's temperature $T_N$, as expected, it
is far superior compared to low temperature performance and the
agreement with Monte-carlo results is even better.
In Fig.\ref{TAP2} we show the results obtained for $K=5$ and $C=10$.
For unbiased messages the system is extremely
sensitive to the choice of initial conditions and does not perform
well in average even at Nishimori's temperature. For biased messages
($f_s=0.1$, $R=1/4$) results are far better and in agreement
with Monte-carlo integration of the RS saddle-point equations.
The experiments show that belief propagation methods may be used
successfully for decoding Sourlas-type codes in practice, and provide
solutions that are well described by RS analytical solutions.
\subsection{Basin of Attraction}
\label{sec:basin}
\begin{figure}
\hspace*{2cm}
\epsfxsize=120mm \epsfbox{figure12.eps}
\caption{Top: Maximum initial deviation $\lambda$ for decoding. Top: $\lambda$
as function of the number of interactions $K$. Circles are
averages over $10$ different codes with $N=300$, $R=1/3$ and noise level
$p=0.1$. Bottom: $\lambda$ as function of the connectivity $C$.
Circles are averages over $10$ codes with $N=300$, $K=3$
and noise level $p=0.1$. Lines and $\times$'s correspond to the RS dynamics
described by the saddle-point equations. }
\label{basins}
\end{figure}
To asses the size of the basin of attraction we consider the decoding process as a dynamics in the graphs space where edges
$\delta q_{\mu j}$ are considered as
dynamical variables. In gauged transformed equations (\ref{eq:gaugetap})
, the perfect decoding of a message correspond to $\delta q_{\mu j}=1$ .
To analyse the basin of attraction we start with random initial values
with a given normalized deviation from the perfect decoding $\lambda=\frac{1}{NC}\sum_{\mu j}(1-\delta q_{\mu j}^{0})$. It is analogous to the finite magnetizations used in the naive mean-field of Section II, since a given $\delta q_{\mu j}^{0}$ corresponds to a given magnetization value by using Eq.(\ref{eq:pseudoposterior}).
In Fig.\ref{basins} we show the maximal deviation in initial conditions required for successful decoding.
Top figure shows an average over $10$ different codes with
$N=300$ (circles) for a fixed code rate $R=1/3$, fixed noise level $p=0.1$ and
increasing $K$. Bottom figure shows the maximal deviation in initial conditions for a fixed number of spins per interaction $K=3$, noise level $p=0.1$ and increasing $C$. We confirm the fidelity of the RS description by comparing the experimental results with the basin of attraction predicted by saddle-point equations (\ref{eq:saddle_point}). One can interpret these
equations as a dynamics in the space of distributions $\pi(x)$.
Performing the transformation $X=\mbox{tanh}(\beta x)$, one can move to
the space of distributions $\Pi(X)$ with support over $[-1,+1]$.
The initial conditions can then be described simply as $\Pi^0(X)=(1-\frac{\lambda}{2})\delta(X-1)+\frac{\lambda}{2}\delta(X+1)$. In Fig.\ref{basins}
we show the basin of attraction of this dynamics as lines and $\times$'s.
The $K=2$ case is the only practical code from a dynamical point
of view, since it has the largest basin of attraction and no prior knowledge on the message is necessary for decoding. Nevertheless, this code's performance degrades faster than the $K>2$ case as shown in Section III, which points to a compromise between good dynamical properties in one side and good performance in the other. One idea could be having a code with changing $K$, starting with $K=2$ to guarantee convergence and progressively increasing its values to improve
the performance \cite{idosaad}.
On the other hand, the basin of attraction increases with $C$. Again it
points to a trade off between good equilibrium properties
(small $C$ and large code rates) and good dynamical properties (large $C$,
large basin of attraction). Mixing small and large $C$ values in the same code
seems to be a way to take advantage of this trade-off \cite{LMSS,davey,VSK}.
\section{Concluding Remarks}
In this paper we studied, using the replica
approach, a finite connectivity many-body spin glass that
corresponds to Sourlas' codes for finite code rates. We have shown,
using a simplified one step RSB solution for
spin glass phase, that for $K\rightarrow\infty$ and $C=\alpha K$
regime at low temperatures the system exhibits a FERRO-SG phase transition
that corresponds to Shannon's bound. However, we have also shown that the
decoding problem for large $K$ has bad convergence properties
when simulated annealing strategies are used.
We were able to find replica symmetric solutions for finite $K$ and
found good agreement with practical decoding performance using belief
networks. Moreover, we have shown that RS saddle-point equations actually
describe the mean behavior of belief propagation algorithms.
We studied the dynamical properties of belief propagation and compared to
statistical physics predictions, confirming the validity of the description. The basin of attraction was shown to depend on $K$ and $C$. Strategies for improving the performance were discussed.
The same methodology has been recently employed successfully \cite{kms99}
to state-of-the-art algorithms as the recent rediscovered Gallager codes
\cite{mackay95} and its variations \cite{idosaad,VSK}.
We believe that the connections found between belief networks and
statistical physics can be further developed to provide deeper insights into
the typical performance of general error-correcting codes.
\label{sec:conclusions}
\ack
This work was partially supported by the program
``Research For The Future'' (RFTF) of the Japanese Society for the
Promotion of Science (YK) and by EPSRC grant GR/L52093 and a Royal Society travel grant (DS and RV).
|
\section{Introduction}
Some five years ago, Seiberg and Witten gave an ansatz for the dominant piece
of the effective action governing the light degrees of freedom of $SU(2)$
${\cal N} = 2$ super Yang-Mills theory at low energy\cite{SeiWitt}. It is
given in terms of an auxiliary complex algebraic curve ${\cal C}$ (whose
moduli space is identified with the quantum moduli space of the low-energy
theory ${\cal M}_\Lambda$) and a given meromorphic differential,
$dS_{SW}$, that induces a special geometry on ${\cal M}_\Lambda$ (see
Refs.\cite{reviews} for excellent reviews).
The solution was soon extended to other gauge groups and matter content by
determining both the appropriate complex curve and meromorphic
differential \cite{sun,klemm,sunmat,hananoz,son}, thus leading to a
substantial progress in our understanding of ${\cal N}=2$ supersymmetric
gauge theories.
In particular, the appearance of an auxiliary Riemann surface made it possible
to identify remarkable connections with string theory, singularity theory of
differentiable maps and integrable systems. The framework that will be used
in this paper is strongly inspired by the latter one.
It is well-known that, as long as ${\cal N}=2$ supersymmetry is
unbroken, the low energy effective action is given in terms of a holomorphic
prepotential ${\cal F}$.
The solution proposed by Seiberg and Witten embodies a prescription to
compute this prepotential.
However, its explicit evaluation for a given gauge group and matter content
is technically involved and it requires
to integrate an expression for the B-periods of $dS_{SW}$ as functions of its
A-periods. The complexity of this procedure increases rapidly with the rank of
the gauge group, even without matter hypermultiplets.
Whereas perturbative contributions to ${\cal F}$ are exhausted by one-loop diagrams
\cite{Sei}, the non-perturbative part is given by an infinite
series of instanton corrections.
The importance of instanton calculus lies precisely in the fact that it
provides one of the few non-perturbative links between the Seiberg--Witten
solution and the microscopic non-abelian field theory that it is supposed to
describe effectively at low energies. From the microscopic point of view, the
first few instanton contributions to the asymptotic semiclassical expansion
of the effective prepotential have been computed for gauge group $SU(N_c)$,
and a remarkable agreement with the Seiberg--Witten solution has been found
\cite{hunter}. From the side of the effective theory, several methods for
determining the instanton corrections have been developed in the last few
years by using the Picard--Fuchs equations \cite{klemm,itoyang1},
holomorphicity arguments \cite{itosas1}, analytic continuation \cite{dhoker}
(also for non-hyperelliptic curves \cite{isa}), modular anomaly equations
\cite{MNW}, etc. Among them, we would like to distinguish those methods that
lead to recursion relations for the $k$-instanton corrections, as long as they
give an implicit expression for the {\em exact} solution.
For the case of $SU(2)$, recursion relations determining the whole instanton
expansion have been obtained both in the pure gauge theory \cite{Matone} and
when matter is included \cite{itoyang1}. These recursion relations were
obtained by combining the renormalization group and Picard-Fuchs equations.
Also one-instanton corrections for the whole ADE series where obtained along
the same lines in \cite{itoyang}.
$SU(N_c)$ with additonal matter in the adjoint representation was considered
in \cite{DHPh} from the point of view of the Calogero-Moser model, and in
\cite{MNW} (for ADE groups with dual Coxeter number $k_D\leq 6$) where the
modular anomaly equations of softly broken ${\cal N}=4$ supersymmetric gauge
theories were invoked.
In a recent paper, \cite{emm}, a new strategy was observed to work
very well for the case of pure $SU(N_c)$.
It is based on a set of first- and second-order equations for the logarithmic
derivatives of the prepotential with respect to the dynamical scale $\Lambda$,
evaluated all over the moduli space
\begin{eqnarray}
\pder{{\cal F}}{\log\Lambda\,} & = & {\beta\over 2\pi i } u_{2}~,
\label{lder} \\
\dpder{{\cal F}}{(\log\Lambda)} & = & -{\beta^2\over 2\pi i}
\pder{u_{2}}{ a^i}\pder{u_{2}}{ a^j}{1\over i\pi}
\partial_{\tau_{ij}}\log\Theta_E(0|\tau) ~, \label{llder}
\end{eqnarray}
where the different quantities entering these expressions will be explained
below. Each one of these equations was obtained separately in the last few
years by many authors \cite{Matone,marimoore,losev,EY,rge}. It was not until
very recently that a unifying approach based on the Whitham hierarchy was
shown to be useful to obtain both \cite{ITEP}. In Ref.\cite{emm}, the ansatz
for the semiclassical expansion of the prepotential for gauge group
$SU(N_c)$ was inserted in both sides of equations (\ref{lder})--(\ref{llder})
with the result of an elegant and systematic procedure that allowed us to
compute instanton corrections {\em up to any desired order} with relatively
little effort. In particular, this method does not require knowledge of the
actual solution for the periods $a$ and $a_D$ of $dS_{SW}$, a fact which
spares a considerable amount of work.
In this paper we will exhibit the strengh of this method by extending the
results of \cite{emm} to any classical gauge group with and without matter
content. We shall limit ourselves to asymptotically free theories. The
obvious question arises, about the validity of equations like (\ref{llder})
for all these situations.
It turns out that, for our present purposes, the only constraint which seems
to be unavoidable within this method is that massive hypermultiplets have to
be introduced in pairs, degenerated in mass.
This can be understood either from a purely theoretical study, which we shall
leave for a separate paper \cite{egrmm}, or else, ``a posteriori", for the
consistency of the results.
Our aim is that this paper could be useful to anyone interested in finding
explicit expressions for the instanton corrections to the effective
prepotential of ${\cal N}=2$ supersymmetric gauge theories. To this end,
general formulas for up to three-instanton corrections will be given and the
method to obtain arbitrary higher corrections will be clearly explained. A
number of particular examples will be also worked out for the lower rank
groups and small number of flavour-pairs, in order to make them easily
available for futher comparison even to other methods or, actually, to the
results obtained in the microscopic non-Abelian field theory.
For the one- and two-instanton corrections, our results coincide with those
in the literature while, for most of the three-instanton contributions, our
results are new.
\section{Recursive Evaluation of the Effective Prepotential}
\setcounter{equation}{0}
\indent
\subsection{ Review and Notation}
For completeness, we shall start by reviewing the r\^ ole of the different
ingredients that enter the formulas (\ref{lder}) and (\ref{llder}).
The low-energy dynamics of ${\cal N}=2$ supersymmetric theories with
classical gauge group $G$ corresponding to $N_c$ colors, and $N_f$
hypermultiplets of mass $m_f = m_{(f+ N_f/2)}$ ({\it i.e.} degenerated in
masses) in the fundamental representation of $G$,
can be described in terms of an auxiliary hyperelliptic curve ${\cal C}$ given
by
\begin{equation}
y^2 = (P(\lambda,e_p)+T(\lambda,m_f,\Lambda))^2 -
4\Lambda^{\beta} F(\lambda,m_f) ~,
\label{curveone}
\end{equation}
where $P$ is the characteristic polynomial of $G$, $\Lambda$ is the quantum
generated dynamical scale, $\beta$ is the coefficient of the one-loop ${\cal
N}=2$ beta function and $e_p$ are the eigenvalues of the complex scalar field
$\langle\phi\rangle = \sum_p e_p h_p$ (see Ref.\cite{Lie} for the conventions
followed in the notation of Lie group and Lie algebra objects) that belongs to
the ${\cal N}=2$ vector supermultiplet in the adjoint of $G$.
$T$ and $F$ are polynomials that do not depend on the {\em moduli} $e_p$ and
$T$ is different from zero only when $N_f>N_c$.
As pointed out in Ref.\cite{dhoker}, when the gauge group is $SU(N_c)$ or
$SO(N_c)$ all dependence on $T$ can be absorbed in a redefinition of
$e_p$, the effective prepotential remaining untouched. Thus,
we can set $T=0$, and write the hyperelliptic curve ${\cal C}$ as
\begin{equation}
y^2 = P^2(\lambda,e_p) - 4\Lambda^{\beta} F(\lambda,m_f) ~.
\label{curve}
\end{equation}
In the case of $Sp(N_c)$, there is a residual value for $T$, $T =
\Lambda^{N_c-N_f+2}(\prod_{f=1}^{N_f/2}m_f^2)$. In order to consider all the
different cases within a unified framework we will neglect this contribution
by setting the mass of one of the degenerated hypermultiplets to zero (say for
example $m_{N_f/ 2}=0$). Following Ref.\cite{dhoker} this case will be
denoted $Sp(N_c)''$, and the corresponding hyperelliptic curve has the form
(\ref{curve}) with the proviso that, according to our previous remark, the
number of hypermultiplets will be at least two, $N_f=N_f'+2\geq 2$.
Consequently, the case of pure $Sp(N_c)$ will not be attainable from our
results though it is quite simple to make the appropriate modifications, as
will be explained below. It is convenient, for later use, to list the form
of $P$, $F$ and $\beta$ for all classical groups:
\vskip7mm
\centerline
{\begin{tabular}{|c||c|c|c|c|c|c|c|c|}
\hline
$G$ & $P(\lambda,e_p)$ & $F(\lambda)$ & $\beta/2$ & $l_G$
& $\xi$ & $\xi^+$ &
$\xi^-$ & $k_D$
\\
\hline\hline
$SU(r+1)$ & $\displaystyle\prod_{p=1}^{r+1}(\lambda-e_p)$ &
$\displaystyle\prod_{f=1}^{N_f}(\lambda+m_f)$ & $r+1-N_f/2$ & r+1 & 1 & 1 & 0
& r+1
\\ \hline
$SO(2r)$ & $\displaystyle\prod_{p=1}^{r}(\lambda^2-e_p^2)$ &
$\displaystyle\lambda^4\prod_{f=1}^{N_f}(\lambda^2-m_f^2)$ & $ 2r-N_f-2$ & r
& 1 & 1 & 1 & 2r-2
\\ \hline
$SO(2r+1)$ & $\displaystyle\prod_{p=1}^{r}(\lambda^2-e_p^2)$ &
$\displaystyle\lambda^2\prod_{f=1}^{N_f}(\lambda^2-m_f^2)$ & $2r-N_f-1$ & r &
1 & 1 & 1 & 2r-1
\\ \hline
$Sp(2r)''$ & $\displaystyle\prod_{p=1}^{r}(\lambda^2-e_p^2)$ &
$\displaystyle\prod_{f=1}^{N_f'}(\lambda^2-m_f^2)$ & $2r-N_f'$ & r & 2 & 2
& 2 & r+1
\\ \hline
\end{tabular}}
\vskip5mm
\centerline{ {\em Table 1}}
\vskip7mm\noindent
The symmetric polynomials $\bar u_k(e_p)$ and $ t_i(m_f)$ are defined through
the expansions
\begin{eqnarray}
P(\lambda,e_p) &\equiv & \lambda^{h} - \sum_{k=2}^{h} \lambda^{h-k}
\bar u_k(e_p) ~,
\label{casimirs} \\
F(\lambda,m_f) &\equiv &\lambda^{h} + \sum_{k=1}^{h} \lambda^{h-k}
t_k(m_f) ~, \label{lastes}
\end{eqnarray}
where $h$ stands, in each case, for the highest power in $P(\lambda,e_p)$ or
$F(\lambda,m_f)$. The moduli for each curve (\ref{curve}) can be taken to be
either the independent roots $e_p$ or the $a^i$ defined as the coefficients
of $\langle \phi\rangle = a^i H_i$ in the Chevalley basis, hence linearly
related to $e_p$ (see Ref.\cite{Lie}). Neither of these parameters are
invariant under Weyl transformations which, in particular, act by permutation
on the $e_p$. On the contrary, the symmetric polynomials (Casimirs) $\bar u_k$
provide faithfull coordinates for the moduli space of vacua.
In particular, $\bar u_2 ={1\over 2}{\rm tr} \phi^2= (\frac{\xi^+
+\xi^-}{2\xi^+})\sum_{p=1}^{l_G} e_p^2 $, and in general:
$\bar u_k = {1\over k} {\rm tr} \phi^k+$ ..., the dots standing for homogeneous
powers of lower Casimir operators.
The Seiberg--Witten meromorphic differential can be written as
\begin{equation}
dS_{SW}(u_k,m_f) = \left({P}' - \frac{P F'}{2 F}\right)
{\lambda d\lambda \over y} ~,
\label{diff}
\end{equation}
and the quantum relations between the low-energy coordinates of the
moduli space $a^i, a_{D\,j}$ and the ``mean field" order parameters $u_k(a) =
{1\over k} {\rm tr} \langle
\phi^k\rangle + \cdots = \bar u_k(a^i) + {\cal O}(\Lambda)$ are implicitely
given by the period integrals
\begin{equation}
a^i(u_k,m_f) = \oint_{A^i} dS_{SW}(u_k,m_f) ~~~~;~~~~
a_{D\,j}(u_k,m_f) = \oint_{B_j} dS_{SW}(u_k,m_f) ~,
\label{theper}
\end{equation}
where $A^i$ and $B_j$ constitute a symplectic basis of homology cycles with
canonical intersections of the hyperelliptic curve (\ref{curve});
the effective prepotential ${\cal F}$ is implicitely defined by the equation
\begin{equation}
a_{D\,i} = \pder{{\cal F}(a)}{a^i} ~,
\label{implicit}
\end{equation}
so that its exact determination involves the integration of
functions $a_{D\,i}(a)$ for which there is not a closed form available.
In this context, the existence of an algorithm that let us determine the
exact form of ${\cal F}$ without going through the actual computation of
$a^i(u_k,m_f)$ and $a_{D\,j}(u_k,m_f)$ is welcome.
As mentioned above, our first ingredient is the set of RG equations
(\ref{lder})--(\ref{llder}). Strictly speaking, as they stand, they are valid
for the pure gauge theory. In presence of matter, the first equation receives
an additional term which depends only on the masses
and the second one has to be modified
for $\beta=2$ where it receives an additional constant contribution
\cite{egrmm}. In summary\footnote{
Although the addition of an $a^i$-independent terms to ${\cal F}$ is unphysical from
the point of view of the effective theory, the embedding of the Seiberg-Witten
solution into the Whitham dynamics fixes them as a function of the bare
coupling $\tau_0$. A similar behaviour was observed in the study of
${\cal F}$ near the strong coupling singularities of $SU(N_c)$
${\cal N}=2$ super Yang--Mills theory \cite{edemas}.}
\begin{eqnarray}
\pder{{\cal F}}{\log\Lambda\,} & = & {\beta\over 2\pi i } u_{2} + \varphi(m)~,
\label{lderm} \\
\dpder{{\cal F}}{(\log\Lambda)\,} & = &
-{\beta^2\over 2\pi i} \pder{u_2}{a^i} \pder{u_2}{a^j} {1\over
i\pi}\partial_{\tau_{ij}}\log\Theta_E(0|\tau) + {\beta^2\over 2\pi i}\Lambda^2
\delta_{\beta,2} ~.
\label{formula}
\end{eqnarray}
In Eq.(\ref{formula}),
$\Theta_E(0|\tau)$ is Riemann's theta function associated to the hyperelliptic
curve ${\cal C}$
\begin{equation}
\Theta \left[ {\vec\alpha\atop \vec\beta}\right]
(\xi|\tau)=
\sum_{n_k \in {\bf Z} }\displaystyle{ e^{i
\pi
\bigl[ \tau_{ij}(n_i+ \alpha_i)
(n_j+ \alpha_j) + 2 (n_i+ \alpha_i)(\xi + \beta_i) \bigr]} }~,
\label{thetaf}
\end{equation}
where $E = \left[ {\vec\alpha\atop \vec\beta}\right]$ stands for an even half
integer characteristic. In almost all the cases this characteristic will be
$E=\left[{0,\dots,0\atop {1\over 2},\dots,{1\over 2}}\right]$, as it is in the case of
pure $SU(N_c)$ \cite{emm}.
The only exception is $Sp(2r)''$ for which the characteristic gets modified to
the value $E=\left[{0,\dots,0,0\atop 0,\dots,0,{1\over 2}}\right]$.
The second ingredient is an ansatz for the instanton expansion of the
prepotential valid for any classical gauge group $G$ and $N_f$ massive
hypermultiplets with paired masses
\begin{eqnarray}
{\cal F}_{G,N_f} & = & \frac{\tau_0^G}{4\pi i}\sum_{\alpha_+} Z_{\alpha_+}^2
+ {i\over 4 \pi}\xi \sum_{\alpha_+} Z_{\alpha_+}^2 \log \,{Z_{\alpha_+}^2\over
\Lambda^2}-{i\over 4\pi}\xi^+
\sum_{p=1}^{l_G}\sum_{f=1}^{N_f/2}(e_p+m_f)^2\log\,{{(e_p+m_f)^2}\over
{\Lambda^2}} \nonumber\\
&& -{i\over 4\pi}\xi^-
\sum_{p=1}^{l_G}\sum_{f=1}^{N_f/2}(e_p-m_f)^2\log\,{{(e_p-m_f)^2}\over
{\Lambda^2}}+{1\over 2 \pi i}\sum_{k=1}^\infty
{\cal F}_{k}(Z)\Lambda^{k\beta} ~,
\label{elprep}
\end{eqnarray}
where $\alpha_+$ denotes a positive root and $\sum_{\alpha_+}$ is the sum
over all positive roots.
The set $\{ \alpha_i \}_{i=1,...,r}$ stands for the simple roots of the
corresponding classical Lie algebra, they generate the root lattice $\Delta =
\{\alpha = n^i \alpha_i| n^i\in {\mathbb{Z}}\}$.
The dot product $(\cdot)$ of two simple roots $\alpha_i$ and $\alpha_j$ gives
an element of the Cartan matrix, $A_{ij}=\alpha_i\!\cdot\!\alpha_j$,
and extends bilinearly to arbitrary linear combinations of simple roots.
So, for example, for any root $\alpha = n^j\alpha_j\in\Delta$,
the quantities $Z_\alpha$ are defined by $Z_\alpha =
a\!\cdot\! \alpha \equiv a^i A_{ij} n^j$ where $a= a^i\alpha_i$. For non-simply
laced Lie algebras this product is not symmetric.
Simple roots can be written in terms of the orthogonal set of unit vectors
$\{\epsilon_p\}_{p=1,\cdots,l_G}$. The order parameters $a^i$ and $e_p$ are related
by $e_p = a\!\cdot\!\epsilon_p$. The exact relations and the actual values of
$\alpha_i\!\cdot\!\epsilon_p$ for each classical gauge group can be found in
Ref.\cite{Lie}. Finally, we also have that $\lambda^k \!\cdot\! \alpha_j=
\delta^k{_j}$ define the fundamental weights. In particular, this means that
$\alpha_i = \sum_k A_{ik}\lambda^k$. We have also introduced three parameters
$\xi,~\xi^+$ and $\xi^-$, so as to deal with all classical Lie algebras within
one single ansatz; $k_D$ denotes the dual Coxeter number. The particular
values of these variables for each classical gauge group are shown in Table 1.
The coefficient of the one-loop beta function turns out to be given by
\begin{equation}
\beta/2 = \xi k_D - \frac{N_f}{2}\left( 1 ~+~ \frac{\xi^-}{\xi^+}\right) ~.
\label{bfun}
\end{equation}
By expressing the roots
$\alpha_+$ in terms of $\epsilon_p$, ({\it i.e.}, $Z_{\alpha_+}$ in
terms of $e_p$) one can also check that
the following relation holds
\begin{equation}
\xi^+~\sum_{p=1}^{l_G} e_p^2 = {1\over k_D}~\sum_{\alpha_+}
Z_{\alpha_+}^2 ~.
\label{rel}
\end{equation}
\subsection{The procedure}
It is important to notice that we may shift $\tau_0^G$ in (\ref{elprep}) to
any value by appropriately rescaling
$\Lambda$, and this will be reflected in our choice for the normalization of
the ${\cal F}_{k}(Z)$.
We have fixed it in all cases to be $\tau_0^G=3\beta/2k_D$ so that quadratic
terms in $Z_\alpha$ do not contribute to the coupling constant $\tau_{ij}$ (see
(\ref{latau}) below).
The l.h.s. of Eq.(\ref{formula}) can be easily computed from the expansion of
the effective prepotential (\ref{elprep}) to be
\begin{equation}
\dpder{{\cal F}}{(\log \Lambda)} = {1\over 2\pi i}
\sum_{k=1}^\infty (k \beta)^2~{\cal F}_k(Z) ~\Lambda^{k\beta} ~,
\label{makdj}
\end{equation}
then comparing (\ref{makdj}) and (\ref{formula}) we get
\begin{equation}
\sum_{k=1}^{\infty}k^2{\cal F}_k\Lambda^{k \beta}=-\pder{u_2}{a^i} \pder{u_2}
{a^j} {1\over i\pi}\partial_{\tau_{ij}}\log \Theta_E(0|\tau)~\label{ic}
+ \Lambda^2 \delta_{\beta,2} ~,
\end{equation}
such that the instanton correction ${\cal F}_k$ can be obtained through a set
of recursive relations after expanding the r.h.s. of Eq.(\ref{ic}) in powers
of $\Lambda^{\beta}$.
The expansion of the derivative of the quadratic Casimir in powers of $\Lambda$
can be obtained from the RG equation (\ref{lderm}) and (\ref{elprep}),
\begin{eqnarray}
\pder{u_2}{a^i} &=& {2\pi i\over \beta} \ppder{{\cal F}}{a^i}{\log\Lambda\,} \cr &=&
{2\xi\over \beta}\sum_{\alpha_+}Z_{\alpha_+}\partial_iZ_{\alpha_+}
-{N_f\over\beta } (\xi^++\xi^-)
\sum_p e_p\partial_i e_p+ \sum_{k=1}^\infty k
{\cal F}_{k,i} ~ \Lambda^{k \beta}\cr
&=& \xi^+\sum_pe_p\partial_ie_p + \sum_{k=1}^\infty k {\cal F}_{k,i} ~
\Lambda^{k \beta}\equiv
\sum_{k=0}^{\infty}{{\cal H}_{2,i}^{(k)}}\Lambda^{k \beta} ~,
\label{expanh}
\end{eqnarray}
where ${\cal F}_{k,i}= \partial{\cal F}_k/\partial a^i $ and use has been made of (\ref{bfun}) and
(\ref{rel}).
To expand the theta function we need to compute from
(\ref{elprep}) the couplings in the semiclassical region,
\begin{eqnarray}
\tau_{ij} &=&\ppder{{\cal F}}{a^i}{a^j} \nonumber \\
&=& {i\over 2\pi}\xi \sum_{\alpha_+} \pder{Z_{\alpha_+}}{a^i}
\pder{Z_{\alpha_+}}{a^j} \log \left({ Z^2_{\alpha_+} \over \Lambda^2} \right)
-{i\over 2\pi}\xi^+
\sum_{p=1}^{l_G}\sum_{f=1}^{N_f/2}\pder{e_p}{a_i}\pder{e_p}{a_j}
\log~{{(e_p+m_f)^2}
\over {\Lambda^2}}\nonumber\\
&&-{i\over 2\pi}\xi^-
\sum_{p=1}^{l_G}\sum_{f=1}^{N_f/2}\pder{e_p}{a_i}\pder{e_p}{a_j}
\log~{{(e_p-m_f)^2} \over {\Lambda^2}}+{1\over 2\pi i} \sum_{k=1}^\infty {\cal F}_{k,ij}
\Lambda^{k \beta} ~,
\label{latau}
\end{eqnarray}
with ${\cal F}_{k,ij}= \ppder{{\cal F}_k}{a^i}{a^j}$. So
the term involving the couplings that appear in the theta function $\Theta_E$
can be written as
\begin{eqnarray}
i\pi \, n^i\tau_{ij} n^j &=& \sum_{\alpha_+} \log
\left({Z_{\alpha_+}\over \Lambda}\right)^{-\xi(\alpha\,\!\cdot\alpha_+)^2}
+~\sum_{p,f}\log~\left(\frac{e_p+m_f}{\Lambda}\right)^{\xi^+(\alpha\,\!\cdot
\epsilon_p)^2}+ \nonumber\\ && + ~\sum_{p,f}
\log~\left(\frac{e_p-m_f}{\Lambda}\right)^{\xi^-(\alpha\,\!\cdot\epsilon_p)^2} +~
{1\over 2}\sum_{k=1}^\infty~ (\alpha\!\cdot\!\!\F''_k\!\!\cdot\! \alpha) ~\Lambda^{k \beta} ~,
\label{tauroot}
\end{eqnarray}
where $\alpha\!\cdot\!\!\F''_k\!\!\cdot\! \alpha\equiv\sum_{i,j}n^i{\cal F}_{k,ij} n^j$.
Also $\alpha= n^i\alpha_i$, and we have set
\begin{equation}
n^i \left(\pder{Z_{\alpha_+}}{a^i}\right) = n^i~(\alpha_i\!\cdot\!\alpha_+) =
\alpha\!\cdot\!\alpha_+ ~~~~~;~~~~~~~ n^i \left(\pder{e_p}{a^i}\right) =
n^i~(\alpha_i\!\cdot\!\epsilon_p) = \alpha\cdot\epsilon_p ~.
\end{equation}
Inserting (\ref{tauroot}) in the Theta function (\ref{thetaf}) with a
characteristic $E=\left[{0,\dots,0\atop {1\over 2},\dots,{1\over 2}}\right]$, we obtain
\begin{eqnarray}
\Theta_E(0|\tau) &=& \sum_{\vec n} \exp\biggl[ i\pi n^i\tau_{ij} n^j + i \pi
\sum_{k} n_k\biggr]
\cr &=&
\sum_{\alpha\in\Delta}(-1)^{\rho\cdot \alpha} \prod_{\alpha_+}\left(
{Z_{\alpha_+}\over \Lambda}\right)^{-\xi(\alpha\,\!\cdot\alpha_+)^2}
\prod_{p,f}\left({{e_p+m_f}\over{\Lambda}}\right)^{\xi^+(\alpha\,\!\cdot\epsilon_p)^2}
\prod_{p,f}\left({{e_p-m_f}\over{\Lambda}}\right)^{\xi^-(\alpha\,\!\cdot\epsilon_p)^2}
\cr &&~~~~~~~\times\prod_{k=1}^\infty\exp\left({{1\over 2} (\alpha\!\cdot\!\!\F''_k\!\!\cdot\! \alpha) \Lambda^{k
\beta}}\right)
\cr &=&
\sum_{s=0}^\infty \sum_{\alpha\in\Delta_s} (-1)^{\rho\cdot\alpha}
\prod_{\alpha_+} Z_{\alpha_+}^{-\xi(\alpha\,\!\cdot\alpha_+)^2}
\prod_{p}\left[ R(e_p)\right]^{(\alpha\,\!\cdot\epsilon_p)^2} \nonumber\\
&&~~~~~~~\times\prod_{k=1}^\infty\left(\sum_{m=0}^\infty {1\over 2^m
m!}\left(\alpha\!\cdot\!\!\F''_k\!\!\cdot\! \alpha\right)^m\, \Lambda^{k\beta m }\right) \Lambda^{s\beta }
\cr &\equiv&
\sum_{l=0}^\infty \Theta^{(l)} \Lambda^{l \beta} ~.
\label{expantheta}
\end{eqnarray}
In the previous expression, $\rho$ is the maximal weight $\rho = \sum_{i=1}^{r}
\lambda^i$. In the case of $Sp(N_c)''$, due to its peculiar characteristic,
the dot product ${\rho\cdot \alpha}$ needs to be replaced by~
${\rho_r\!\cdot\alpha= n^r}$. On the other hand, $\Delta_s\subset\Delta$ is a
subset of the root lattice composed of those lattice vectors
$\alpha\in\Delta_s$ that fulfill the constraint
$
\frac{1}{2}\xi^+ \sum_p(\alpha\!\cdot \epsilon_p)^2 = s ~.
$
The Weyl group permutes the $\epsilon_p$, hence the previous statement is Weyl
invariant. In other words, the sets $\Delta_s$ are unions of Weyl orbits.
This fact guarantees that the final result will recombine into Weyl invariant
expressions. In (\ref{expantheta}) we have also introduced the polynomial
\begin{eqnarray}
R(\lambda,m_f) &=& \prod_{f=1}^{N_f/2}(\lambda+m_f)^{\xi^+}
\prod_{f=1}^{N_f/2}(\lambda-m_f)^{\xi^-} \label{erres}\\
&\equiv& \left(\lambda^h + \sum_{i=1}^{h}
q_i(m_f)\lambda^{h-i}\right)^{\xi^+} ~,
\label{lasquses}
\end{eqnarray}
where $h$, again, stands for the appropriate highest power. It is at this point
where the need for pairwise equal masses enters. Otherwise, we would be
dealing with square root factors of the form $(\lambda
\pm m_f)^{1/2}$. Now we can collect the first few terms in
the expansion (\ref{expantheta}),
\[ \Theta^{(0)} = 1 ~, ~~~~~~~~~~
\Theta^{(1)} = \sum_{\alpha\in\Delta_1}(-1)^{\rho\cdot\alpha}
\prod_{\alpha^+} Z_{\alpha_+}^{-\xi(\alpha\cdot\alpha_+)^2}
\prod_p\left[R(e_p)\right]^{(\alpha\cdot \epsilon_p)^2} ~, \]
\begin{eqnarray}
\Theta^{(2)} & = & \sum_{\alpha\in\Delta_1}(-1)^{\rho\cdot\alpha}\,{1\over 2}
(\alpha\!\cdot\!\!\F''_1\!\!\cdot\! \alpha ) \prod_{\alpha^+} Z_{\alpha_+}^{-\xi(\alpha\cdot\alpha_+)^2}
\prod_p\left[R(e_p)\right]^{(\alpha\cdot \epsilon_p)^2} \nonumber\\
& & + \sum_{\beta\in\Delta_2}(-1)^{\rho\cdot\beta} \prod_{\alpha^+}
Z_{\alpha_+}^{-\xi(\beta\cdot\alpha_+)^2}
\prod_p\left[R(e_p)\right]^{(\alpha\cdot \epsilon_p)^2} \nonumber ~.
\end{eqnarray}
However, in the logarithmic derivative, the Theta function appears in the
denominator so we shall need the expansion of $\Theta(0|\tau)^{-1}$ in terms of
$\Lambda$. We can write this expansion as
\begin{equation}
\Theta(0|\tau)^{-1} = \sum_{l=0}^{\infty} \Xi^{(l)}(\Theta) \,
\Lambda^{2Nl} ~.
\label{tauinv}
\end{equation}
Here $\Xi^{(0)}(\Theta) = 1$ and for $\Xi^{(l)}(\Theta)$ we can write
in general
\begin{equation}
\Xi^{(l)}(\Theta) = \sum_{(p_1,...,p_k) \in {\bf N}^k}^{p_1 + 2p_2 + ...
+ kp_k = l} \chi_{(p_1,...,p_k)} \prod_{i=1}^l (\Theta^{(i)})^{p_i} ~,
\label{tauinvcoef}
\end{equation}
where the coefficients $\chi$ are parametrized by the
partition elements $(p_1,...,p_k)$. The first few values for these
parameters are, for example,
\[ \chi_{(1)} = -1 ~, ~~~ \chi_{(2,0)} = 1 ~, ~~~ \chi_{(0,1)} = -1 ~, ~~~
\chi_{(3,0,0)} = -1 ~, ~~~ \chi_{(1,1,0)} = 2 ~, ~~~ \chi_{(0,0,1)} = -1 ~, \]
and using these values we can immediately obtain the lower $\Xi^{(l)}(\Theta)$.
Next, we compute the derivative of
the theta function with respect to the period matrix
\begin{eqnarray}
{1\over i\pi}\partial_{\tau_{ij}}\Theta_E(0,\tau) & = & \sum_{ n} n^i n^j
\exp\biggl[ i\pi n^k\tau_{kl} n^l + i \pi \sum_{k} n_k\biggr] \cr
& = & \sum_{s=1}^\infty \sum_{\alpha\in\Delta_s} (-1)^{\rho\cdot\alpha}
(\lambda^i\!\cdot\alpha)(\lambda^j\!\cdot\alpha)
\prod_{\alpha_+} Z_{\alpha_+}^{-\xi(\alpha\,\!\cdot\alpha_+)^2}
\prod_{p}\left[ R(e_p)\right]^{(\alpha\,\!\cdot\epsilon_p)^2} \nonumber\\
& & ~~~~~~\times\prod_{k=1}^\infty\exp\left({{1\over 2} (\alpha\!\cdot\!\!\F''_k\!\!\cdot\! \alpha)
\Lambda^{k\beta }}\right) \Lambda^{s \beta}\nonumber\\
&\equiv& \sum_{p=1}^\infty \Theta_{ij}^{(p)} \Lambda^{p\beta} ~.
\label{expanthij}
\end{eqnarray}
Now, collecting all the pieces and inserting them back into (\ref{formula}),
we find for ${\cal F}_k(Z)$ the following expression:
\begin{equation}
{\cal F}_k(Z) = - k^{-2}
\sum_{p, q, l=0}^{p+q+l = k-1}\sum_{ij} {\cal H}_{2,i}^{(p)} {\cal
H}_{2,j}^{(q)} \Theta_{ij}^{(k-p-q-l)} \Xi^{(l)} ~,
\label{elresul}
\end{equation}
in terms of previously defined coefficients.
If we look at the factors on the r.h.s. of Eq.(\ref{elresul}), it is easy
to see that they involve ${\cal F}_1,{\cal F}_2,...$ up to
${\cal F}_{k-1}$. In fact, although both ${\cal H}_2^{(p)}$ and $\Theta^{(p)}$ depend
on ${\cal F}_1,....{\cal F}_p$, the indices within parenthesis reach at most the value
$k-1$ as $\Theta_{ij}^{(0)}=0$. Moreover $\Theta_{ij}^{(k)}$ depends on
${\cal F}_1,...,{\cal F}_{k-1}$ since the vector $\alpha=0$ is missing from the lattice
sum. This ``lucky accident" has its origin in the particular form of
the characteristic in the semiclassical (duality) frame, and seems to be an
essential feature in order to build up a recursive procedure to {\em compute
all the instanton coefficients by starting just from the perturbative
contribution to} ${\cal F}(a)$ \cite{edemas}. For the first few cases we may develop
(\ref{elresul}) to find
\begin{eqnarray}
{\cal F}_1 &=& -
\sum_{ij} {\cal H}^{(0)}_{2,i} {\cal H}^{(0)}_{2,j} \Theta_{ij}^{(1)} \nonumber
\\
{\cal F}_2 & = & - \frac{1}{4} \sum_{ij}
\left(\Theta_{ij}^{(2)} {\cal H}_{2,i}^{(0)} {\cal H}_{2,j}^{(0)}
+ \Theta_{ij}^{(1)}(2 {\cal H}_{2,i}^{(1)}{\cal H}_{2,j}^{(0)} -
{\cal H}_{2,i}^{(0)}{\cal H}_{2,j}^{(0)}\Theta^{(1)})\right) \nonumber\\
&\vdots& \nonumber
\end{eqnarray}
After some algebraic manipulations (\ref{elresul}) admits the following
general form
\begin{equation}
{\cal F}_k=-\frac{1}{k^2} \sum_{s=1}^k\sum_{ \alpha\in\Delta_{s}}
(-1)^{\rho\cdot\alpha}
\prod_{\alpha_+} Z_{\alpha_+}^{ -\xi (\alpha\,\!\cdot\alpha_+)^2}
\prod_{p}\left[R(e_p)\right]^{ (\alpha\,\!\cdot\epsilon_p)^2}
\Phi_{k+1-s}(\alpha) ~,
\label{ff}
\end{equation}
where the functions $\Phi_{k} (\alpha)$ depend on ${\cal F}_1\cdots, {\cal F}_{k-1}$ and
have to be evaluated case by case.
For the first few we have
\begin{eqnarray}
\Phi_1(\alpha) &= & Z_{\alpha(G)}^2 ~, \label{phi0} \\
\Phi_2({\cal F}_1,\alpha) & = & {\cal F}_1 + 2 (\alpha\!\cdot\!\!\F'_1 ) Z_{\alpha(G)} + {1\over 2} (\alpha\!\cdot\!\!\F''_1\!\!\cdot\! \alpha)
Z_{\alpha(G)}^2 ~, \label{phi1} \\
\Phi_3({\cal F}_1,{\cal F}_2,\alpha) & = & 4{\cal F}_2 + 4 (\alpha\!\cdot\!\!\F'_2) Z_{\alpha(G)} + (\alpha\!\cdot\!\!\F'_1)^2
+ {1\over 2} (\alpha\!\cdot\!\!\F''_1\!\!\cdot\! \alpha)\left( {\cal F}_1 + 2 (\alpha\!\cdot\!\!\F'_1) Z_{\alpha(G)} \right) \nonumber \\
& & + {1\over 8} (\alpha\!\cdot\!\!\F''_1\!\!\cdot\! \alpha)^2 Z_{\alpha(G)}^2 + {1\over 2} (\alpha\!\cdot\!\!\F''_2\!\!\cdot\! \alpha) \, Z_{\alpha(G)}^2 ~,
\label{phi2}
\end{eqnarray}
where $\alpha\!\cdot\!\!{\cal F}'_k = n^i{\cal F}_{k,i}$.
Expressions (\ref{ff})-(\ref{phi2}) make patent the iterative character of
the procedure. ~$Z_{\alpha(G)}$ stands for $n^i {\cal H}_{2,i}^{(0)}$~, and for
simply laced groups, $Z_{\alpha(G)}=Z_\alpha$ while for non-simply laced, the
exact form will be given below. In the case $\beta =2$, as we can see from
Eq.(\ref{ic}) the first instanton correction acquires a shift
$\tilde{{\cal F}}_1={\cal F}_1+1~.$ It soon becomes clear that, except for the simplest
cases, the concrete evaluation of the ${\cal F}_k$ has to be carried out by
symbolic computation. In the next sections we illustrate our procedure with
explicit examples in several cases for the lower rank groups.
A last word concerning the possibility to split the masses is in order.
Generically, all resulting expressions involve powers of the degenerated masses
$\{m_f,~f=1,...,N_f/2\}$. They must be recovered from the exact result for
arbitrary masses in the coincidence limit $m_f = m_{f+N_f}$. The possibility
to go back unambiguously only happens for low powers of $m_f$. As a thumb
rule, we have checked in several cases that the following prescription does
the job: for $SU(N_c)$ and when powers of $m_f$ are not higher than 2,
$m_f\to {1\over 2}(m_f + m_{f+N_f})$ and $m_f^2 \to m_f m_{f+N_f}$; while for the
rest of the groups and for (even) powers of $m_f$ not higher than 4,
$m_f^2\to {1\over 2}(m_f^2 + m^2_{f+N_f})$ and $m_f^4 \to m^2_f m^2_{f+N_f}$.
Typically these cases only occur in ${\cal F}_1$. Another way to see this is to
observe that if we write down ${\cal F}_1$ in terms of the $q_i(m_f),~
f=1,..N_f/2$ as given in (\ref{lasquses}), these factors always appear
precisely in those combinations that build up the $t_k= t_k(q_i)$ as given in
(\ref{lastes}), which are valid for arbitrary masses.
\section{Results for Simply-Laced Lie Algebras}
\setcounter{equation}{0}
\indent
We start by giving the concrete expression for $Z_{\alpha(G)} =
n^i{\cal H}_{2,i}^{(0)}$. In the case of $A_r$ and $D_r$ Lie algebras, we have that
$\sum_p e_p^2 = a^i a^j A_{ij} $ and also that $\xi^+=1$. Hence
$$
Z_{\alpha(G)} = n^i{\cal H}_{2,i}^0 =n^i\sum_p e_p\partial_i e_p=n^i a^j A_{ij} = a\!\cdot\!
\alpha = Z_\alpha
$$
We shall define $\Delta_0=\prod_{\alpha_+}Z_{\alpha+}^2~. $ To avoid
confusion we must mention that, although we have expressed all results in
terms of Weyl invariant polynomial combinations $\bar u_k(a^i)$, for
notational clearness we shall drop the bar.
\subsection{{\boldmath $SU(r+1)$} with {\boldmath $N_f$} hypermultiplets}
The only asymptotically free theories that we can consider within our approach,
for these groups, are $N_f=2,\ldots,2r$. Let us list some of the results that
we have obtained by using our formulas (we omit the case of $SU(N_c)$ whithout
matter which can be found in \cite{emm}):
\subsubsection{\boldmath $SU(2)$}
\noindent\underline{$N_f=2$}~
Using $u_2=a^2$, we found the following corrections:
\begin{eqnarray}
{\cal F}_1 &=& \rule{0mm}{10mm}\frac{u_2+m^2}{2u_2} ~,
\label{su2f1}\\
{\cal F}_2 &=& \rule{0mm}{10mm}\frac{u^2_2-6u_2 m^2 + 5 m^4}{64 u^3_2} ~,
\label{su2f2} \\
{\cal F}_3 &=& \rule{0mm}{10mm}\frac{5u^2_2 m^2-14u_2 m^4+9m^6}{192u_2^5} ~.
\label{su2f3}
\end{eqnarray}
The one- and two-instanton contributions coincide with those computed in
\cite{dhoker} (after adjusting $\Lambda$ to $\Lambda/2$).
In the one-instanton correction, as discussed before, it is possible to
split the masses, $m^2 \to m_1m_2$, so that the result for non-degenerated
$(ND)$ matter hypermultiplets is
\begin{equation}
{\cal F}_{1,ND} = \rule{0mm}{10mm}\frac{u_2+m_1m_2}{2u_2} ~,
\label{su2non}
\end{equation}
in agreement with the result in Ref.\cite{dhoker}.
Also, the case of one hypermultiplet can then be considered by letting
$m_2\to\infty$ while keeping $\Lambda m_2$ finite an equal to the square of
the dynamical scale $\Lambda^2$ that corresponds to $N_f=1$.
\subsubsection{\boldmath $SU(3)$}
To express our results, we introduce Weyl invariant combinations
in terms of the
$a_i$-variables, $u_2=a_1^2+a_2^2-a_1a_2$ and $u_3=a_1a_2 (a_1-a_2)$.
For the case $N_f=4$, we will denote $q_1=m_1+m_2$ and $q_2=m_1m_2$. We
obtained:
~
\noindent\underline{$N_f=2$}
\begin{eqnarray}
{\cal F}_1 &=&\rule{0mm}{10mm} {(2u_2^2+6m^2u_2-18mu_3)}/{\Delta_0} ~,
\label{su3f1} \\
{\cal F}_2 &=&\rule{0mm}{10mm}
(5u_2^6+153m^4u_2^4+162m^2u_2^5-1998m^3u_3u_2^3-414mu_3 u_2^4+1701m^4u_3^2u_2
\nonumber\\ &&+4374m^2u_3^2u_2^2+162u_3^2u_2^3+729u_3^4-2916m^3u_3^3-2673
mu_3^3u_2)/{\Delta_0^3} ~,
\label{su3f2} \\
{\cal F}_3 &=&\rule{0mm}{10mm}(48u_2^{10}+12320m^6u_2^7+31792m^4u_2^8+4992m^2u_2^9
-366624m^5u_3u_2^6-12032mu_3u_2^8\nonumber\\
&&-253088m^3u_3u_2^7+478116m^6u_3^2u_2^4+2276856m^4u_3^2
u_2^5+529236m^2u_3^2u_2^6\nonumber\\
&&+5600u_3^2u_2^7-3684852m^5u_3^3u_2^3-4654800m^3u_3^3u_2^4
-394524mu_3^3u_2^5+994356m^6u_3^4u_2\nonumber\\
&&+7097544m^4u_3^4u_2^2+3969648m^2u_3^4u_2^3+105192u_3^4u_2^4
-1469664m^5u_3^5\nonumber\\
&&-4878468m^3u_3^5u_2-1571724mu_3^5u_2^2+1364688m^2u_3^6+
215784u_3^6u_2)/{\Delta_0^{5}} ~.
\label{su3f3}
\end{eqnarray}
Again, the splitting of the one-instanton correction can be done by letting
$m^2 \to m_1m_2$ and $m \to (m_1+m_2)/2$,
\begin{equation}
{\cal F}_{1,ND} = \rule{0mm}{8mm}(2u_2^2+6m_1m_2u_2-9(m_1+m_2)u_3)/{\Delta_0} ~,
\label{su3non1}
\end{equation}
and the reduction of hypermultiplets mentioned above is immediate.
Our results (\ref{su3f1}), (\ref{su3f2}) and (\ref{su3non1}) agree with those
obtained in \cite{dhoker}.
~
\noindent\underline{$N_f=4$}
\begin{equation}
{\cal F}_1 = \rule{0mm}{8mm}(2u_2^3-9u_3^2-6q_1u_2u_3+2(q_1^2+2q_2)u_2^2+6q_2^2u_2-
18q_1q_2u_3)/{\Delta_0} ~,
\label{su3bf1}
\end{equation}
and the length of the expressions growths rapidly.
Again, having in mind that ${\cal F}_{1,ND}$ should be linear in the polynomials of
the masses it is possible to carefully split the masses, $m_1^2 \to m_1m_3$,
$m_2^{\,2} \to m_2m_4$, $m_1 \to (m_1+m_3)/2$ and $m_2 \to (m_2+m_4)/2$ (or, in
terms of the polynomials of the masses, $2q_1 \to t_1$, $q_1^2+2q_2 \to t_2$,
$2q_1q_2 \to t_3$ and $q_2^2 \to t_4$) so that
\begin{equation}
{\cal F}_{1,ND} = \rule{0mm}{8mm}(2u_2^3-9u_3^2-3t_1u_2u_3+2t_2u_2^2-9t_3u_3
+6t_4u_2)/{\Delta_0} ~,
\label{su3non2}
\end{equation}
and then one can reduce it to an odd number of matter hypermultiplets
in the way mentioned before \cite{dhoker}.
\subsubsection{\boldmath $SU(4)$}
As the expressions become too long, we will only display the
one-instanton correction. To express our results,
we introduce the classical
values of the Weyl invariant Casimirs
$u_2=a_1^2+a_2^2+a_3^2-a_1a_2-a_2a_3$,
$u_3=a_1a_2(a_1-a_2)+a_2a_3(a_2-a_3)$ and $u_4=a_1^2a_2a_3-a_1a_2^2a_3
-a_1^2a_3^2+a_1a_2a_3^2$.
For the case $N_f=4$, we will denote $q_1=m_1+m_2$ and $q_2=m_1m_2$.
\vskip2mm
\noindent\underline{$N_f=2$}
\begin{eqnarray}
{\cal F}_1 &=& (8m^2u_2^3+6u_3^2u_2-8mu_3u_2^2-36u_3^2m^2+32m^2u_2u_4
\nonumber\\
&& -16u_2^2u_4+96mu_3u_4-64u_4^2)/{\Delta_0} ~.
\label{su4f1}
\end{eqnarray}
\noindent\underline{$N_f=4$}
\begin{eqnarray}
{\cal F}_1&=&(2u_3^2u_2^2-8u_4u_2^3+12u_3^2u_4-32u_4^2u_2-18q_1u_3^3
+64q_1u_4u_3u_2\nonumber\\
&&+6(q_1^2+2q_2)u_3^2u_2-16(q_1^2+2q_2)u_4u_2^2-64(q_1^2+2q_2)
u_4^2-8q_1q_2u_3u_2^2\nonumber\\
&&+96q_1q_2u_4u_3+8q_2^2u_2^3-36q_2^2u_3^2+32q_2^2u_4u_2)/
{\Delta_0} ~.
\label{su4bf1}
\end{eqnarray}
\subsection{{\boldmath $SO(2r)$} with {\boldmath $N_f$} hypermultiplets}
The only asymptotically free theories that we can consider within our
approach, for these groups, are $N_f=0,2,\ldots,2(r-2)$. Notice that the case
$N_f=0$ corresponds to vanishing values of $\xi^+$ and $\xi^-$ in
(\ref{elprep}), which in the formulas of the instanton corrections implies
$R=1$. Let us list some of the result one can easily obtain by using our
formulas:
\subsubsection{\boldmath $SO(4)$}
For this group the classical values of the Casimir operators in terms of the
$a_i$ are given by $u_2=2a_1^2+2a_2^2$, $u_4=-(a_1^4+a_2^4-2a_1^2a_2^2)$, and
we can only consider the pure case ($N_f=0$).
~
\noindent\underline{$N_f=0$}
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{5mm}2^2u_2/{\Delta_0} ~, \\
{\cal F}_2&=&\rule{0mm}{10mm}2(5u_2^3-60u_2u_4)/{\Delta_0^3} ~, \\
{\cal F}_3&=&\rule{0mm}{10mm}2^5(3u_2^5-120u_2^3u_4+240u_4^2u_2)/{\Delta_0^5} ~.
\end{eqnarray}
\subsubsection{\boldmath $SO(6)$}
For this group we have $u_2=2a_1^2+2a_2^2+2a_3^2-2a_1a_2-2a_1a_3$,
$u_4=-(a_1^4+a_2^4+a_3^4-2a_1^3a_2-2a_1a_2^3-2a_1^3a_3-2a_1a_3^3+3a_1^2a_2^2
+3a_1^2a_3^2-2a_2^2a_3^2-2a_1^2a_2a_3+2a_1a_2^2a_3+2a_1a_2a_3^2)$ ,
$u_6=a_1^4a_2^2-2a_1^3a_2^3+a_1^2a_2^4-2a_1^4a_2a_3+2a_1^3a_2^2a_3+a_1^4a_3^2
+2a_1^3a_2a_3^2-2a_1^2a_2^2a_3^2-2a_1^3a_3^3+a_1^2a_3^4$,
and we can consider the cases $N_f=0$ and $N_f=2$.
~
\noindent\underline{$N_f=0$}
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{5mm}2^2(-u_2u_4-9u_6)/{\Delta_0} ~,\\
{\cal F}_2&=&\rule{0mm}{10mm}2(-5u_2^5u_4^3-43u_2^3u_4^4-172u_2u_4^5-
60u_2^6u_4u_6-647u_2^4u_4^2u_6-1701u_2^2u_4^3u_6+36u_2^5u_6^2
\nonumber\\
&&-1827u_2^3u_4u_6^2-3276u_4^4
u_6+1323u_2u_4^2u_6^2-14337u_2^2u_6^3-21627u_4u_6^3)/{\Delta_0^3} ~.
\end{eqnarray}
\noindent\underline{$N_f=2$}
\begin{equation}
{\cal F}_1=\rule{0mm}{5mm}2^2(-m^4u_2u_4-9m^4u_6-4m^2u_4^2+u_4^2u_2+12m^2u_2u_6
-4u_6u_2^2-3u_4u_6)/{\Delta_0} ~.
\end{equation}
Again, the splitting of the masses is possible for the one-instanton
correction. This happens for all the classical groups.
\subsubsection{\boldmath $SO(8)$}
In this case the expressions of the Casimir operators in terms of the $a_i$
gets too long so we are not going to list them here. We have found the
following expressions:
\noindent\underline{$N_f=0$}
\begin{equation}
{\cal F}_1=2^2(9u_2^3u_8-u_2^2u_4u_6+3u_2u_6^2+32u_2u_4u_8+
48u_6u_8-4u_4^2u_6)/{\Delta_0} ~.
\end{equation}
\noindent\underline{$N_f=2$}
\begin{eqnarray}
{\cal F}_1&=&2^2(9m^4u_2^3u_8-m^4u_2^2u_4u_6+3m^4u_2u_6^2+32m^4u_2
u_4u_8+48m^4u_6u_8-4m^4u_4^2u_6\nonumber\\ &&
-4m^2u_6^2u_2^2-12m^2u_4u_6^2-u_2u_4u_6^2-9u_6^3+12m^2u_4u_8u_2^2+32m^2u_8u_4^2
+4u_8u_4^2u_2\nonumber\\
&&+16m^2u_8u_6u_2-3u_8u_6u_2^2+32u_8u_6u_4+128m^2u_8^2-48u_8^2u_2)/{\Delta_0}
~.
\end{eqnarray}
The one-instanton corrections agree with those computed in Ref.\cite{dhoker}.
\section{Results for non Simply-Laced Lie Algebras}
\setcounter{equation}{0}
\indent
\subsection{{\boldmath $SO(2r+1)$} with {\boldmath $N_f$} hypermultiplets}
In this case the form of $Z_{\alpha(G)}$ is different from the simply-laced
cases. Indeed, using the fact that
\[ \alpha_i\!\cdot\!\epsilon_q = \delta_{i,q}-\delta_{i+1,q}
~~~~~~~~~~~ \alpha_r\!\cdot\!\epsilon_q = 2\delta_{r,q} ~, \]
and setting $\xi^+=1$, we have
\begin{eqnarray}
Z_{\alpha(G)} &=& n^i{\cal H}_{2,i}^{(0)} ~= ~n^i\sum_pe_p\partial_ie_p
\nonumber\\& = &n^i\sum_pe_p(\alpha_i\!\cdot\epsilon_p)
= \sum_{i=1}^{r-1}~\left(e_i-e_{i+1}\right)n^i+2 e_r n^r
\nonumber\\ & = & Z_\alpha + Z_{\alpha_r} n^r ~.
\label{zetagso}
\end{eqnarray}
The only asymptotically free theories that we can consider within our approach,
for $SO(2r+1)$, are $N_f=0,2,\ldots,2r-2$.
Notice that the case $N_f=0$ corresponds, as in $SO(2r)$,
to take in (\ref{elprep}) $\xi^+=\xi^-=0$ which in the formulas
of the instanton corrections means to set $R=1$. Let us list some
of the results that we have obtained:
\subsubsection{\boldmath $SO(5)$}
We can consider within our approach the cases $N_f=0$ and $N_f=2$. For this
group we have $u_2=2a_1^2+4a_2^2-4a_1a_2$,
$u_4=-(a_1^4-4a_1^3a_2+4a_1^2a_2^2)$. We found:
~
\noindent\underline{$N_f=0$}
\begin{eqnarray}
{\cal F}_1 &=& \rule{0mm}{5mm}-2^3u_4/{\Delta_0} ~,
\label{so5f1}
\\
{\cal F}_2 &=&\rule{0mm}{10mm} 2(u_2^3u_4^2-76u_2u_4^3)/{\Delta_0^3} ~,
\label{so5f2}
\\
{\cal F}_3 &=&
\rule{0mm}{10mm}2^{7}(3u_2^4u_4^4-232u_2^2u_4^5+176u_4^6)/{3\Delta_0^5} ~.
\label{so5f3}
\end{eqnarray}
\noindent\underline{$N_f=2$}
\begin{eqnarray}
{\cal F}_1 &=&\rule{0mm}{6mm} 2^2 (-2m^4u_4+2m^2u_2u_4-u_2^2u_4-2u_4^2)/{\Delta_0} ~,
\label{so5bf1}
\\
{\cal F}_2&=&\rule{0mm}{10mm}2(m^8u_2^3u_4^2-76m^8u_2u_4^3+152m^6u_2^2u_4^3
-32m^6u_4^4 -78m^4u_2^3u_4^3+168m^4u_2u_4^4\nonumber\\
&&+12m^2u_2^4u_4^3-88m^2u_2^2u_4^4+96m^2u_4^5-u_2^5u_4^3
-60u_2u_4^5+u_2^3u_4^4)/{\Delta_0^3} ~,
\label{so5bf2}
\\
{\cal F}_3 &=& \rule{0mm}{10mm}2^6(6m^{12}u_2^4u_4^4-464m^{12}u_2^2u_4^5
+352m^{12}u_4^6-9m^{10}u_2^5u_4^4+1256m^{10}u_2^3u_4^5\nonumber\\
&&-1744m^{10}u_2u_4^6+3504m^8u_2^2u_4^6+3m^8u_2^6u_4^4-1212m^8u_2^4u_4^5
-960m^8u_4^7\nonumber\\
&&+2976m^6u_2u_4^7-3024m^6u_2^3u_4^6+498m^6u_2^5u_4^5-2864m^4u_2^2u_4^7
+1054m^4u_2^4u_4^6\nonumber\\
&&-86m^4u_2^6u_4^5+736m^4u_4^8-1104m^2u_2u_4^8+824m^2u_2^3u_4^7
-137m^2u_2^5u_4^6\nonumber\\
&&+5m^2u_2^7u_4^5+240u_2^2u_4^8-128u_4^9-56u_2^4u_4^7+5u_2^6u_4^6)/{3\Delta_0^5}
~.
\end{eqnarray}
\subsubsection{\boldmath $SO(7)$}
Here we have $u_2=2a_1^2+2a_2^2+4a_3^2-2a_1a_2-4a_2a_3$~, $u_4=
-(a_1^4+a_2^4-2a_1^3a_2-2a_1a_2^3-4a_2^3a_3+3a_1^2a_2^2+8a_1^2a_3^2+4a_2^2a_3^2
-8a_1^2a_2a_3+8a_1a_2^2a_3-8a_1a_2a_3^2)$,
$u_6=a_1^4a_2^2+a_1^2a_2^4-2a_1^3a_2^3-4a_1^4a_2a_3+8a_1^3a_2^2a_3
-4a_1^2a_2^3a_3+4a_1^4a_3^2-8a_1^3a_2a_3^2+4a_1^2a_2^2a_3^2$~, and for $N_f=4$
we also denote $q_2=m_1^2+m_2^{2}$ and $q_4=m_1^2m_2^{\,2}$. We can consider
the cases $N_f=0$, $N_f=2$ and $N_f=4$.
~
\noindent\underline{$N_f=0$}
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{5mm}2^3(u_2^2u_6+3u_4u_6)/{\Delta_0} ~,
\\
{\cal F}_2&=&\rule{0mm}{10mm}2\, u_6^2(-u_2^6u_4^3-12u_2^4u_4^4-48u_2^2u_4^5-
64u_4^6-76u_2^7u_4u_6\nonumber\\
&&-839u_2^5u_4^2u_6-3057u_2^3u_4^3u_6-3588u_2u_4^4u_6-44u_2^6u_6^2
-1695u_2^4u_4u_6^2\nonumber\\
&&-10827u_2^2u_4^2u_6^2-16308u_4^3u_6^2+1863u_2^3u_6^3-567u_2u_4u_6^3
-18225u_6^4)/{\Delta_0^3} ~.
\end{eqnarray}
\noindent\underline{$N_f=2$}
\begin{equation}
{\cal F}_1=2^3u_6(m^4u_2^2+3m^4u_4+m^2u_2u_4+9m^2u_6-3u_2u_6)/{\Delta_0} ~.
\end{equation}
\noindent\underline{$N_f=4$}
\begin{eqnarray}
{\cal F}_1&=&2^3u_6(q_4^2u_2^2+3q_4^2u_4-3(q_2^2+2q_4)u_2u_6
+(q_2^2+2q_4)u_4^2+q_2q_4u_2u_4+9q_2q_4u_6+3q_2u_4u_6\nonumber\\
&&+4q_2u_2^2u_6-q_2u_2u_4^2-45u_6^2-32u_2u_4u_6
-8u_2^3u_6+7u_4^3+2u_2^2u_4^2)/{\Delta_0} ~.
\end{eqnarray}
Again, the one-instanton corrections agree with previous results \cite{dhoker}.
\subsection{{\boldmath $Sp(2r)$} with {\boldmath $N_f$} hypermultiplets}
In this subsection we are going to consider the case of $Sp(2r)''$,
{\it i.e.}, the case of $Sp(2r)$ with two
massless hypermultiplets and $N_f'=N_f-2$ matter hypermultiplets.
In this case the form of $Z_{\alpha(G)}$ is, as in the case of $SO(2r+1)$,
different from the simply-laced cases. Now, using
\[ \alpha_i\!\cdot\!\epsilon_q = \delta_{i,q}-\delta_{i+1,q}
~~~~~~~~~~~~~~~~ \alpha_r\!\cdot\!\epsilon_q = \delta_{r,q} ~, \]
we see that
\begin{eqnarray}
Z_{\alpha(G)} &=& n^i{\cal H}_{2,i}^{(0)}\nonumber\\
& \equiv & 2n^i\sum_pe_p\partial_ie_p
=2 n^i \sum_p e_p(\alpha_i\!\cdot\epsilon_p)
= 2\left[\sum_{i=1}^{r-1}\left(e_i-e_{i+1}\right)n^i+ e_r
n^r\right] \nonumber\\ & = & 2 Z_{\alpha} - Z_{\alpha_r}n^r ~.
\label{zetagsp}
\end{eqnarray}
The only asymptotically free theories that we can consider within our approach,
for $Sp(2r)''$, are $N_f'=0,2,\ldots,2r-2$.
Notice that the case $N_f'=0$ now means to put $R(e_p)=e_p^4$ cause
we are considering two massless hypermultiplets. Let us list some
of the results that we have obtained:
\subsubsection{\boldmath $Sp(4)$}
For this group we can consider the cases $N_f'=0$ and $N_f'=2$. We have
$u_2=2a_1^2+a_2^2-2a_1a_2$, $u_4= -(a_1^4-2a_1^3a_2+a_1^2a_2^2)$.
~
\noindent\underline{$N_f'=0$}
\begin{eqnarray}
{\cal F}_1&=\rule{0mm}{5mm}&u_2/{2\Delta_0} ~,
\\
{\cal F}_2&=&\rule{0mm}{10mm}(5u_2^5+43u_2^3u_4+172u_2u_4^2)/4{\Delta_0^3} ~,
\\
{\cal F}_3&=&\rule{0mm}{10mm}2^2(9u_2^9+143u_2^7u_4+927u_2^5u_4^2+2840u_2^3u_4^3
+5680u_2u_4^4)/3{\Delta_0^5} ~.
\end{eqnarray}
\noindent\underline{$N_f'=2$}
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{5mm}(m^4u_2+4m^2u_4-u_2u_4)/2{\Delta_0} ~,
\\
{\cal F}_2&=&\rule{0mm}{10mm}(5m^8u_2^5+43m^8u_2^3u_4+172m^8u_2u_4^2+12m^6u_2^4u_4
-24m^6u_2^2u_4^2+352m^6u_4^3+54m^4u_2^3u_4^2\nonumber\\
&&-264m^4u_2u_4^3+152m^2u_2^2u_4^3-32m^2u_4^4-5u_2^3u_4^3
+60u_2u_4^4)/4{\Delta_0^3} ~,
\\
{\cal F}_3&=&\rule{0mm}{10mm}2^2(9m^{12}u_2^9+143m^{12}u_2^7u_4+927m^{12}u_2^5u_4^2
+2840m^{12}u_2^3u_4^3\nonumber\\
&&+5680m^{12}u_2u_4^4+28m^{10}u_2^8u_4+304m^{10}u_2^6u_4^2+1536m^{10}u_2^4u_4^3
-1408m^{10}u_2^2u_4^4\nonumber\\
&&+9728m^{10}u_4^5-11952m^8u_2u_4^5+6600m^8u_2^3u_4^4+75m^8u_2^7u_4^2
+537m^8u_2^5u_4^3\nonumber\\
&&+15872m^6u_2^2u_4^5-1200m^6u_2^4u_4^4+120m^6u_2^6u_4^3-1792m^6u_4^6
+7760m^4u_2u_4^6\nonumber\\
&&+321m^4u_2^5u_4^4-5416m^4u_2^3u_4^5-3584m^2u_2^2u_4^6+1280m^2u_4^7
+752m^2u_2^4u_4^5\nonumber\\
&&-9u_2^5u_4^5+360u_2^3u_4^6-720u_2u_4^7)/3{\Delta_0^5} ~.
\end{eqnarray}
\subsubsection{\boldmath $Sp(6)$}
For this group we can consider the cases $N_f'=0$, $N_f'=2$ and $N_f'=4$.
We let $u_2=2a_1^2+2a_2^2+a_3^2-2a_1a_2-2a_2a_3$, $u_4=
-(a_1^4+a_2^4-2a_1^3a_2-2a_2^3a_3-2a_1a_2^3+3a_1^2a_2^2+2a_1^2a_3^2+a_2^2a_3^2
-4a_1^2a_2a_3+4a_1a_2^2a_3-2a_1a_2a_3^2)$ ,
$u_6=a_1^4a_2^2-2a_1^3a_2^3+a_1^2a_2^4-2a_1^4a_2a_3+4a_1^3a_2^2a_3
-2a_1^2a_2^3a_3+a_1^4a_3^2-2a_1^3a_2a_3^2+a_1^2a_2^2a_3^2$. For $N_f'=4$, and
we also have $q_2=m_1^2+m_2^{\,2}$ and
$q_4=m_1^2m_2^{2}$.
~
\noindent\underline{$N_f'=0$}
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{5mm}2(3u_2u_6-u_2^2u_4-4u_4^2)/{\Delta_0} ~,
\\
{\cal F}_2&=&\rule{0mm}{10mm}2^4(-5u_2^6u_4^5-60u_2^4u_4^6-240u_2^2u_4^7-320u_4^8
+43u_2^7u_4^3u_6+546u_2^5u_4^4u_6+2304u_2^3u_4^5u_6\nonumber\\
&&+3232u_2u_4^6u_6-172u_2^8u_4u_6^2-2255u_2^6u_4^2u_6^2
-9715u_2^4u_4^3u_6^2-13272u_2^2u_4^4u_6^2+2064u_4^5u_6^2\nonumber\\
&&+180u_2^7u_6^3-1107u_2^5u_4u_6^3-18975u_2^3u_4^2u_6^3-46908u_2u_4^3u_6^3
+2439u_2^4u_6^4\nonumber\\
&&-3240u_2^2u_4u_6^4-57672u_4^2u_6^4+19197u_2u_6^5)/{\Delta_0^3} ~.
\end{eqnarray}
\noindent\underline{$N_f'=2$}
\begin{eqnarray}
{\cal F}_1=\rule{0mm}{7mm}2(3m^4u_2u_6-m^4u_2^2u_4-4m^4u_4^2-4m^2u_2^2u_6
-12m^2u_4u_6-u_2u_4u_6-9u_6^2)/{\Delta_0} ~.
\end{eqnarray}
\noindent\underline{$N_f'=4$}
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{7mm}2(3q_4^2u_2u_6-q_4^2u_2^2u_4-4q_4^2u_4^2-4q_2q_4u_2^2u_6
-12q_2q_4u_4u_6-(q_2^2+2q_4) u_2u_4u_6\nonumber\\
&&-9(q_2^2+2q_4)u_6^2+12q_2u_2u_6^2-4q_2u_4^2u_6-3u_4u_6^2-4u_2^2u_6^2
+u_2u_4^2u_6)/{\Delta_0} ~.
\end{eqnarray}
\subsection{The case of pure {\boldmath $Sp(2r)$}}
As we discussed above, the case of $Sp(2r)$ without matter hypermultiplets
cannot be obtained from our previous formulas, as long as we are
considering at least two massless hypermultiplets.
Nevertheless, one can treat this case separately in an analogous way.
In fact, we can fix our ansatz for the effective prepotential (\ref{elprep}) to
the one first considered by Ito and Sasakura \cite{itosak} by setting
\begin{equation}
\xi=1~~~~~~~~~\xi^+=\xi^-=0 ~~~~~~~~\mbox{and}~~~~~~~~~~ \tau_0 = 3 ~.
\label{settings}
\end{equation}
Now, we can introduce the effective prepotential into Eq.(\ref{formula}) and
the same kind of formulas for the instanton correction would be obtained,
provided we have for this case a characteristic $E=\left[{0,\dots,0\atop
{1\over 2},\dots,{1\over 2}}\right]$. Note that, being $N_f=0$, we must set $R=1$ in our
formulas.
We also need the value of $Z_{\alpha(G)}$ which turns out to be the same
as that in $Sp(2r)''$, \ie \ $Z_{\alpha(G)}=2Z_\alpha-n_rZ_{\alpha_r}$.
\subsubsection{\boldmath $Sp(4)$}
For this group we have, as we saw before, $u_2=2a_1^2+a_2^2-2a_1a_2$, $u_4=
-(a_1^4-2a_1^3a_2+a_1^2a_2^2)$. In terms of them the first instanton
corrections are
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{5mm}2^3(u_2^2+4u_4)/{\Delta_0} ~,
\\
{\cal F}_2&=&\rule{0mm}{10mm}2^6(5u_2^7+59u_2^5u_4+232u_2^3u_4^2
+304u_2u_4^3)/{\Delta_0^3} ~,
\\
{\cal F}_3&=&\rule{0mm}{10mm}2^{14}(9u_2^{12}+184u_2^{10}u_4+1526u_2^8u_4^2
+6496u_2^6u_4^3+14656u_2^4u_4^4\nonumber\\
&&+15872u_2^2u_4^5+5632u_4^6)/{3\Delta_0^5} ~.
\end{eqnarray}
\subsubsection{\boldmath $Sp(6)$}
For this group we have $u_2=2a_1^2+2a_2^2+a_3^2-2a_1a_2-2a_2a_3$, $u_4=
-(a_1^4+a_2^4-2a_1^3a_2-2a_2^3a_3-2a_1a_2^3+3a_1^2a_2^2+2a_1^2a_3^2+a_2^2a_3^2
-4a_1^2a_2a_3+4a_1a_2^2a_3-2a_1a_2a_3^2)$ ,
$u_6=a_1^4a_2^2-2a_1^3a_2^3+a_1^2a_2^4-2a_1^4a_2a_3+4a_1^3a_2^2a_3
-2a_1^2a_2^3a_3+a_1^4a_3^2-2a_1^3a_2a_3^2+a_1^2a_2^2a_3^2$, and the first
instanton corrections are
\begin{eqnarray}
{\cal F}_1&=&\rule{0mm}{5mm}2^5(u_2^2u_4^2+4u_4^3-4u_2^3u_6-18u_2u_4u_6
-27u_6^2)/{\Delta_0} ~,
\\
{\cal F}_2&=&\rule{0mm}{10mm}
2^{12}(-5u_2^6u_4^7-60u_2^4u_4^8-240u_2^2u_4^9-320u_4^{10}+59u_2^7u_4^5u_6
+738u_2^5u_4^6u_6\nonumber\\
&&+3072u_2^3u_4^7u_6+4256u_2u_4^8u_6-232u_2^8u_4^3u_6^2-3021u_2^6u_4^4u_6^2
-12699u_2^4u_4^5u_6^2\nonumber\\
&&-15736u_2^2u_4^6u_6^2+6288u_4^7u_6^2+304u_2^9u_4u_6^3+4120u_2^7u_4^2u_6^3
+15518u_2^5u_4^3u_6^3+1716u_2^3u_4^4u_6^3\nonumber\\
&&-55728u_2u_4^5u_6^3-16u_2^8u_6^4+5928u_2^6u_4u_6^4+54486u_2^4u_4^2u_6^4
+113373u_2^2u_4^3u_6^4-216u_2^5u_6^5\nonumber\\
&&-41148u_4^4u_6^4+39447u_2^3u_4u_6^5+182250u_2u_4^2u_6^5-729u_2^2u_6^6
+89667u_4u_6^6)/{\Delta_0^3} ~.
\end{eqnarray}
The one-instanton corrections agree with those computed in Ref.\cite{itosak}.
\section{Concluding Remarks}
\setcounter{equation}{0}
\indent
In the present paper, we have shown how instanton corrections to the effective
prepotential of ${\cal N}=2$ supersymmetric theories can be computed in a
variety of cases including all classical gauge groups and even number of
degenerated fundamental matter hypermultiplets, {\em up to arbitrary order}.
As compared to other approaches developed in the literature, we should stress
that the one presented in this paper has an important feature in that it does
not require an explicit knowledge of the BPS spectrum as a function of the
moduli, at the same time that it allows to consider a huge variety of
cases within a unified framework.
Also, being recursive, it admits an easy implementation on a computer.
We have illustrated the remarkable simplicity of our procedure by displaying
many explicit expressions which should be quite useful for further comparison
with the results obtained by other means.
Conversely, our results admit a second reading: They could be thought of as a
highly non-trivial test of the connection between the Seiberg--Witten solution
to the low energy dynamics of ${\cal N}=2$ supersymmetric gauge theories, and
the theory of Whitham (adiabatic) deformations of a given integrable system
\cite{emm,ITEP}. In this sense, it is important to remind that the {\em new}
equation (\ref{formula}), which is a key ingredient of our procedure, is
originated in the latter framework, as it is shown in detail in
Ref.\cite{egrmm}.
Aside from being an interesting mathematical problem by itself, the embedding
of the Seiberg-Witten solution within a Whitham hierarchy seems to be the
appropriate framework for the study of many physical phenomena.
For example, the so-called {\em slow Whitham times} can be consistently
thought of as spurion vector supermultiplets that can be used to break ${\cal
N}=2$ supersymmetry down to ${\cal N}=0$ with non-quadratic Casimir operators
\cite{emm}.
In this way, the Whitham hierarchy can be interpreted as a family of
supersymmetry breaking deformations of the original theory associated with the
higher Casimir operators of the gauge group.
This issue generalizes to the ${\cal N}=0$ case the family of ${\cal N}=1$
supersymmetry breaking terms considered, for instance, in Ref.\cite{ad}.
The key feature of the Whitham formalism lies on the fact that, as the
dependence on the slow times is encoded in the prepotential, it is possible to
obtain the {\em exact} effective potential of the theory, in the spirit of
\cite{emm,soft}.
This allows to perform a detailed study, both qualitative and quantitative,
of the vacuum state of the theory once supersymmetry is broken, as well as
of the appearance of monopole condensates, mass gaps, etc.
Preliminary results on this program were published in Ref.\cite{emm}. The
formalism is even useful near the Argyres--Douglas singularities, where
non-local degrees of freedom become simultaneously massless, provided one
approaches them along any of the submanifolds where a unique monopole gets
massless \cite{emm}.
There is another place where deformations of the prepotential by means of the
Whitham times are relevant. It is the study of contact terms in the twisted
version of ${\cal N}=2$ gauge theories, where these new variables play the
r\^ole of sources for insertions of certain class of operators in the
generating functional \cite{marimoore,losev,wm,takasaki}.
Another interesting point is given by the uses of our starting equations
(\ref{lderm})--(\ref{formula}) to study the strong coupling expansion of the
prepotential near the singularities of the quantum moduli space, as it was
done in Ref.\cite{edemas} for the case of pure $SU(N_c)$.
In particular, these equations provide us with a set of non-trivial
constraints that facilitate the study of the couplings between different {\em
magnetic photons}, originally found in Ref.\cite{ds}, that take place at such
points. The expansion of the prepotential near the maximal points by other
methods, as the deformations of the auxiliary singular Riemann manifold
\cite{dHP}, is not sensitive to such kind of terms.
Several interesting questions remain open aside from the ones just mentioned.
For example, the case of arbitrary masses cannot be treated within our
approach, except for the one-instanton correction (which, on the other hand,
is enough for leading order comparison purposes). In fact, from the Whitham
hierarchy side, one can show that indeed the formulas used in this paper are
insufficient to tackle the generic scenario, though it seems to be possible to
refine the formalism in order to extend its applicability to some cases of
unpaired masses \cite{egrmm}. The additional corrections that appear in the
generic case are, nevertheless, quite difficult to manage with. Finally,
another avenue for future research is, certainly, the connection of this
formalism with the string theory and D-brane approach to supersymmetric gauge
theories, where some steps has already been given in the last few years
\cite{gorsky}.
We believe that these matters deserve further study.
\section*{Acknowledgements}
We are pleased to thank Marcos Mari\~no for helpful comments and a careful
reading of the manuscript. The work of J.D.E. is supported by a fellowship of
the Ministry of Education and Culture of Spain. The work of J.M. was
partially supported by DGCIYT under contract PB96-0960.
|
\section{Introduction}
The possibility that massive neutron stars might be driven unstable to collapse
to black holes when placed in a close binary orbit was first suggested
by Wilson, Mathews and Maronetti (hereafter WMM; ~\cite{WWM}) on the basis
of their approximate relativistic numerical simulations.
This finding was quite unexpected, partly because it disagreed
with earlier Newtonian calculations ~\cite{LRS} which showed that tidal fields
stabilize binary stars against radial collapse. In fact, none of the follow-up
post-Newtonian (PN)~\cite{LAI2} ~\cite{LOM} or approximate
analytic analyses ~\cite{Kip} indicate
the presence of any relativistic radial instability in fluid binaries,
nor does an independent dynamical simulation ~\cite{Shib}.
No evidence is found for the WMM ``crushing effect'' in the relativistic
numerical calculations of either corotational ~\cite{Baum} or irrotational ~\cite{Bon}
binaries in quasi-equilibrium circular orbits. These later calculations are particularly
careful to adopt the same simplifications as WMM (e.g, a conformally flat 3-metric).
All of the calculations except those of WMM suggest that the maximum allowed rest mass
of a fluid star in a binary is in fact slightly larger than the value in isolation.
While it would appear that the WMM effect may not be present for fluid stars,
the work of WMM raises the interesting question as to whether
the ``crushing instability'' might exist for a different type of binary
system. Specifically, are there {\it any} relativistic binary systems for which tidal fields
can trigger the collapse of a compact object known to be stable in isolation?
To address this question, Shapiro ~\cite{Us} offered a simple candidate compact object
consisting of a test particle orbiting a Schwarzschild black
hole outside the innermost-stable circular orbit (ISCO). Such a ``compact object''
is obviously stable in isolation, but when it is placed in a binary
orbit about a distant mass, an integration of the test-particle equations of motion reveals that
the tidal field of the distant mass can cause a test particle to plunge inside the hole.
Shapiro's model is very promising as a simple illustration of the crushing instability
in action, but the original analysis, as he emphasized, was highly simplified and somewhat heuristic.
First, the study was confined to two special test-particle orbits, one coplanar with and the other
perpendicular
to the companion orbital plane. Moreover, the motion of the particle in the perpendicular case
was artificially constrained to remain in a plane, precluding any precessional motion or wobbling.
Most important, the analysis was based on a post-Newtonian treatment of the 3-body
problem in which the tidal piece of the equation for
the relative motion of the test particle about the black hole was treated to
lowest (Newtonian) order and the nontidal piece replaced by the exact, fully relativistic expression
for geodesic motion in Schwarzschild geometry. While such a hybrid approach, similar in spirit to one
proposed by Kidder, Will and
Wiseman~\cite{LEK} for the 2-body problem, should capture the essential dynamics, it is
far from rigorous. In particular, it is not clear {\it a priori}
how reliable it is to use a Newtonian tidal term
when the test particle is moving at high velocity in a very strong gravitational field close
to a black hole.
In this paper we improve and extend the simple model presented in ~\cite{Us}
of a compact object subject to binary-induced collapse.
We use the equations of Regge and Wheeler ~\cite{RW} and Zerilli ~\cite{Z}
to derive the perturbation to the
spacetime close to a Schwarzschild black hole due to a distant mass. We note in
passing that
the resulting perturbed spacetime provides another example supporting the recent
conjecture of Bekenstein ~\cite{Bek} that the
horizon area of a perturbed black hole is an adiabatic invariant. We obtain the geodesic
equations of motion in the perturbed spacetime and solve them to track the dynamical evolution
of a dense spherical swarm of 20,000 test particles placed around the hole outside the
ISCO. The `black hole $+$ swarm' constitutes
a compact collisionless cluster whose stability against tidally-induced collapse we determine.
We compare our refined treatment with the simpler version presented in ~\cite{Us} and show that
the original equations track the behavior reasonably well.
We confirm that the crushing instability can occur in a binary system containing simple compact
collisionless clusters like the ones we construct.
\section{HYBRID -- PN TREATMENT}
\subsection{Basic Equations}
Consider first a system of three bodies. Assume that two of the
bodies, 1 and 2, say, are much closer to each other than they are to the third, so that
the influence of body 3 on the relative orbit of 1 and 2 may be treated as a small
tidal perturbation.
Specialize to the case in which body 1 is a Schwarzschild
black hole of mass $M$ and body 2 is a test particle ($m_t \ll M $).
Since it is far away from the 1-2 pair (the ``compact object''), body 3 can be
treated as a point particle, with mass $m$.
Let {\bf r} be the coordinate position of the test particle relative to the
hole and {\bf R} be the coordinate position of the distant point mass relative to the
hole. Following Shapiro ~\cite{Us}, we may write the equation for the relative motion
of the test particle about the hole as
\begin{equation}
\label{six}
\ddot{\bf r} =\ -\ \left({M \over r^2}\right)\ [A{\bf e_r} + B{\bf v}]
+ {m \over R^3} \ \left[ {3{\bf R} \cdot {\bf r} \over R^2}
\ \ {\bf R}\ \ -\ \ {\bf r} \right]\, .
\end{equation}
where ${\bf e_r} = {\partial\over\partial r}$.
In the Newtonian limit,
$A=1$ and $B=0$ in Eq.~(\ref{six}). For
isolated binaries, Lincoln and Will~\cite{CWL} derive
post-Newtonian expressions for $A$ and $B$ for arbitrary
masses, correct through 2.5PN order. Kidder, Will and
Wiseman~\cite{LEK} provide a ``hybrid'' set
of equations in which the sum of the terms in $A$ and $B$
that are independent of the ratio
$\eta = \mu /(M +m_t), \mu = M m_t/(M +m_t) ,$ is replaced by the exact
expression for geodesic motion in the Schwarzschild geometry around
a body of mass $M$, while the terms dependent on
$\eta$ are left unaffected. Their resulting equation of
motion is therefore exact in the test-body limit $(\eta
\rightarrow 0)$ and is valid to 2.5PN order when
appropriately expanded for arbitrary masses. We shall
utilize these same hybrid expressions for $A$ and $B$ and,
for simplicity, work in the test-body limit by taking one
member of our close pair to have a mass much smaller than
the other. Adopting harmonic (de Donder) coordinates, the
resulting (Schwarzschild) expressions for $A$ and $B$ are
given by
\begin{equation}
\label{twelve}
A \ =\ \left[{1-M /r\over (1\ +\ M /r)^3}
\right]\ \ -\ \
\left[ {2-M /r \over 1-(M /r)^2}\right]\ {M \over r}\
\dot{r}^2\ +\ v^2 ,\end{equation}
\begin{equation}
\label{thirteen}
B\ =\ -\ \left[{4-2M /r \over
1-(M /r)^2}\right]\ \dot{r}\, ,
\end{equation}
where
\begin{equation}
\label{fourteen}
v^2\ =\ \dot{r}^2\ +\ r^2 (\dot{\theta}^2 + \sin^2\theta\dot{\phi}^2).
\end{equation}
To model the three-body system, we must also know the
position of the companion (body 3) relative to the compact object.
The leading Newtonian piece of the equation of motion for {\bf R}
gives
\begin{equation}
\label{five}
\ddot {\bf R}\ =\ -{M+m \over R^3}{\bf R} \, .
\end{equation}
We are interested in the dynamical behavior of the close
pair, regarded as a single ``compact object'', as it
inspirals toward the distant mass $m$ due to gravitational
radiation emission. To treat the inspiral of this
``binary'' ($m$ in orbit about the ``compact object''),
we must include radiation reaction terms in the lowest order
(Newtonian) orbit Eq.~(\ref{five}). Formally, such a
treatment requires a consistent expansion up to 2.5PN
order. In lieu of this, we shall analyze the inspiral by
assuming that the binary is in a nearly circular, Keplerian
orbit, which undergoes a slow inspiral due to gravitational
radiation loss in the quadrupole limit. This assumption is
equivalent to inserting a quadrupole radiation reaction
potential in the binary orbit Eq.~(\ref{five}) and neglecting the
lower-order, (non-dissipative) PN corrections in that
particular equation. While such an expression is not
formally consistent to 2.5PN order, it
faithfully tracks the secular inspiral of the binary in the
limit treated here in which the binary system is at wide
(nonrelativistic) separation. The details of the inspiral
are not important here, only that the inspiral serves to
bring a tidal perturber slowly in from infinity toward our
compact object~\cite{OUR}. The resulting equations for the
binary inspiral are then ~\cite{Us}
\begin{equation}
\label{fifteen}
R(t)/R(0) = (1 -t/T)^{1/4}
\end{equation} and
\begin{equation}
\label{sixteen}
\theta(t) -\theta(0) ={8 \over 5}\left({ M + m
\over R(0)^3}
\right)^{1/2}T [1 - (1-t/T)^{5/8}] \, ,
\end{equation}
(Note the typo in ~\cite{Us}.) The binary inspiral timescale $T$ is given by
\begin{equation}
\label{seventeen}
T/M = {5 \over 256} {(R(0)/M )^4
\over {((m+M)/M)(m/M)}} \, .
\end{equation}
We will use Eqs.~(\ref{fifteen})--(\ref{seventeen}) in analyzing the relative orbit
Eq.~(\ref{six}).
\subsection{Numerical Implementation}
To assess the fate of the compact object, we followed the motion of
a spherical swarm of 20,000 test particles about the black hole.
At $t=0$, the test particles were placed randomly about the hole at a radius $r/M=5.9$,
well outside the ISCO of an isolated black hole at $r/M = 5$, and set in circular
orbits of arbitrary orientation. We set $m = M$ and started
the simulation with the companion at $R/M = 60$. At this initial
separation, the tidal field of the companion is negligible at the compact object, and the
test particles begin in stable equilibrium orbits. We integrate (\ref{six}) until
the time at which $R/M = 26.30$. Cartesian coordinates were used throughout.
The results of the simulation are summarized in Figure 1, where we plot the mean cluster
radius as a function of companion separation.
Clearly, the compact object
remains stable until the companion comes within $R/M \approx 30$, at which point the
tidal field causes many of the particles to plunge into the hole.
We find that
17,919 particles (89.6 \%) have fallen into the black hole by $t/M = 121,893$, the
time at which the companion reaches $R/M = 26.30$.
The simulation confirms the existence of a crushing instability.
\begin{figure}
\epsfxsize=3in
\begin{center}
\leavevmode
\epsffile{fig1a.ps}
\end{center}
\epsfxsize=3in
\begin{center}
\leavevmode
\epsffile{fig1b.ps}
\end{center}
\caption{Evolution of the cluster in the hybrid-PN approximation.
The mean radius of the 20,000 test-particles in the initially spherical swarm
is plotted as a function of the tidal companion radius in the top figure and
as a function of coordinate time in the bottom figure.}
\end{figure}
Several checks were performed to ensure the accuracy of our code, which
integrates the ordinary differential orbit
equations by a standard fourth-order Runge Kutta algorithm with an adaptive stepsize ~\cite{NR}.
For example, in the absence of the companion, the code
conserves particle energy and angular momentum over the full time period of integration
and reliably locates the Schwarzschild ISCO with a controllable precision. These tests are
not so trivial when the integrations are performed in Cartesian coordinates.
\section{Schwarzschild Perturbation Treatment}
\subsection{Basic Equations}
\begin{figure}
\epsfxsize=1.5in
\begin{center}
\leavevmode
\epsffile{fig2.ps}
\end{center}
\caption{The adopted coordinate system. The black hole $M$ is at the origin and the
companion $m$ is on the $z$ axis.}
\end{figure}
We will now improve the previous treatment by incorporating the tidal field of the
companion star in a fully relativistic fashion. To do this, we treat the effect of the companion as a
small perturbation on a Schwarzschild background. By assuming that the companion orbits at a much
greater distance from the black hole than the test particles, and hence moves with a much smaller angular
velocity, we can work in the quasi-static approximation. We
place the companion on the $z$ axis of our
spherical polar coordinate system centered on the black hole (see Figure 2).
We follow Regge and Wheeler~\cite{RW} and divide the linear metric
perturbations into independent ``even'' and ``odd'' components
according to
\begin{equation}
\label{2one}
g_{\mu\nu}\ =\ g_{\mu\nu}(S) + h_{\mu\nu}(\mbox{odd}) + h_{\mu\nu}(\mbox{even}) \, ,
\end{equation}
where $g_{\mu\nu}(S)$ is the usual Schwarzschild metric.
In the appropriate gauge, Regge and Wheeler found that the odd perturbations may be written as
\begin{equation}
\label{2two}
\begin{array}{l}
\mathflushleft{h_{\mu\nu}(\mbox{odd})\ =\ \pmatrix{
0 & 0 & 0 & h_0(r) \cr
0 & 0 & 0 & h_1(r) \cr
0 & 0 & 0 & 0 \cr
h_0(r) & h_1(r) & 0 & 0 \cr}\times}\cr
\mathstrutter{15pt} \mathflushright{\sin\theta {\partial \over \partial\theta}Y_{L0}(\theta)\times
e^{(i\omega t)} \, .}\cr
\end{array}
\end{equation}
Likewise, the even perturbations may be written, after simplifying gauge transformations, as
\begin{equation}
\label{2three}
\begin{array}{l}
\mathflushleft{h_{\mu\nu}(\mbox{even}) =}\\
\mathstrutter{35pt}\mathcenter{\pmatrix{\scriptstyle H_0(r)(1-2M/r) &\scriptstyle H_1(r) &\scriptstyle 0 &\scriptstyle 0 \cr
\scriptstyle H_1(r) & \mathstrutter{12pt}{\scriptstyle H_2(r) \over (1-2M/r)} & \scriptstyle 0 & \scriptstyle 0 \cr
\scriptstyle 0 & \scriptstyle 0 & \scriptstyle r^2 K(r) & \scriptstyle 0 \cr
\scriptstyle 0 & \scriptstyle 0 & \scriptstyle 0 & \scriptstyle r^2 \sin^2\theta K(r) \cr
}\times}\\
\mathstrutter{15pt}\mathflushright{Y_{L0}(\theta)\times e^{(i\omega t)} \, .}\\
\end{array}
\end{equation}
We now proceed
to solve Einstein's equations to first order in the perturbation
functions as in Zerilli ~\cite{Z}, using
a delta function (point) source for the companion. We first find
solutions to the homogeneous equations away from the companion
and then use the inhomogeneous source terms to match these solutions.
Setting $\omega = 0$ to comply with our quasi-static approximation, Zerilli's equations
reduce to
\begin{equation}
\label{2four}
h_1 = H_1 = 0; \ \ \ \ \ H_0\ =\ H_2 \equiv H \,
\end{equation}
\begin{equation}
\label{2five}
\begin{array}{l}
\displaystyle
\mathflushleft{{d^2 h_0 \over dr^2} + \left[{4M \over r^2}-{L(L+1) \over r}\right]{h_0 \over r-2M} = }\\
\mathstrutter{15pt}\mathflushright{-8\pi r\left[{1 \over 2}L(L+1)\right]^{-1/2}Q_{L}^{(0)}\, ,}\\
\end{array}
\end{equation}
\begin{equation}
\label{2six}
\begin{array}{l}
\mathflushleft{\bigl(1 - {2M \over r}\bigr)^2 {d^2 K \over dr^2} + \bigl(1-{2M \over r}\bigr)\bigl(3-
{5M \over r}\bigr){1 \over r}{dK \over dr}}\\
\mathstrutter{15pt}\mathonefourth{- \bigl(1-{2M \over r}\bigr)^2\ {1 \over r}\ {dH \over dr}
- \bigl(1-{2M \over r}\bigr){1 \over r^2}(H-K)}\\
\mathstrutter{15pt}\mathflushright{- \bigl(1-{2M \over r}\bigr) {1 \over 2r^2}L(L+1)(H+K) = 8\pi A_L^{(0)} \, ,}\\
\end{array}
\end{equation}
\begin{equation}
\label{2seven}
\mathcenter{-{1-M/r \over r-2M}{dK \over dr}+{1 \over r}{dH \over dr} + {2-L(L+1) \over 2r(r-2M)}(H-K) = 8\pi A_L \, .}
\end{equation}
In the above equations, $Q_L$, $A_L$, and $A_L^{(0)}$ represent components of the stress-energy tensor.
For a point source companion moving along a Schwarzschild geodesic, these terms are given in Appendix E of
Zerilli ~\cite{Z}. From
Zerilli's tabulation of the source terms, one sees that all the source terms go to zero in the
static limit except for $A_L^{(0)}$, i.e. $T_{00}$.
Homogeneous solutions may be found analytically for both the even and odd perturbations. Regge and
Wheeler show that, in the large $r$ limit (where the companion resides), there is one
solution for $h_0$ varying like $r^{-L}$, and another varying like $r^{L+1}$. Since there is no
source term for $h_0$ ($Q_L = 0$), the only way that both $h_0$ and
its first derivative can be continuous and regular everywhere is to require
$h_0 = 0$. Therefore, there is no contribution from odd parity
perturbations.
The static solutions for $H$ and $K$ are found
both by Regge and Wheeler~\cite{RW} and by Zerilli~\cite{Z} in terms of hypergeometric functions.
We are interested in the lowest order nontrivial contribution, the quadrupole piece,
generated by the companion, so
we restrict our attention to the $L=2$ perturbation. Inside the orbit of the companion,
the solution which is regular at the horizon is
\begin{equation}
\label{2eight}
\begin{array}{ll}
\displaystyle
H \ =\ \kappa_1 r(r-2M) & \approx\kappa_1 r^2 \\
\displaystyle
K \ =\ \kappa_1 (r^2-2M^2) & \approx\kappa_1 r^2 \, ,
\end{array}
\end{equation}
where the last equality in each equation above holds at large $r \gg M$ and $\kappa_1$ is
a constant to be determined by
matching at the companion. These equations may be verified by direct substitution into
Eq.~(\ref{2six}) or (\ref{2seven}). Outside the orbit of the companion, the perturbation is
given by the solution regular at infinity,
\begin{equation}
\label{2nine}
\begin{array}{ll}
H\ &=\ \kappa_2 \Bigl[{2M (2M^3+4M^2r-9M r^2+3r^3) \over (2M -r)r} \\
& \hskip90pt +{3r^2(r-2M )^2\ln(1-2M /r) \over (2M -r)r}\Bigr] \\
& \approx \kappa_2 r^{-3} \\
& \\
K\ &=\ \kappa_2 M^2\bigl[-6+{{4M}\over{r}}-{{6r}\over{M}}-{{3r^2\ln(1-2M /r)}\over{M^2}} \\
& \mathstrutter{15pt} \hskip 120pt + 6\ln(1-2M /r)\bigr] \\
& \approx \kappa_2 r^{-3} \, ,
\end{array}
\end{equation}
where the last equality in each equation again holds at large $r \gg M$ and $\kappa_2$ is
another constant to be determined by matching.
Now we determine the two constants by matching the solutions at the radius of
orbit of the companion, $R$.
Because $A_L^{(0)}$, the source term of (\ref{2six}), does not go
to zero in the quasi-static case, $\kappa_1$ and $\kappa_2$ do not vanish, as was the case for
the odd parity solutions.
Taking $R \gg M$ as required by the quasi-static approximation,
we use the asymptotic expressions in equations (\ref{2eight}) and (\ref{2nine}) for the matching.
Requiring that $K$ be continuous across $R$ then yields our first condition
\begin{equation}
\label{2ten}
\kappa_1 R^2 = \kappa_2 R^{-3} ~~\Rightarrow~~ \kappa_2 = R^5 \kappa_1 \, .
\end{equation}
The second condition may be found in either of two ways. First, one may integrate Equation (\ref{2six})
across the source to obtain a
jump condition relating difference in the first derivative of $K$ on either side of $r=R$ to the
source strength. Alternatively, one may compute $\kappa_1$ directly by taking the Newtonian
limit of $H$ and matching it to the Newtonian tidal potential. Either way, one obtains
$\kappa_1 = {{m}\over{R^3}} 4\sqrt{\pi/5}$. Thus, $h_{00}$, for example, is
\begin{equation}
\label{2thirteen}
\begin{array}{l}
\mathflushleft{h_{00} = (1-2M /r)(r^2-2M r)\kappa_1Y_{20}(\theta)}\\
\mathstrutter{15pt}\mathflushright{= (1-2M /r)(r^2 - 2M r){m \over R^3}(3\cos^2\theta - 1)\, .}\\
\end{array}
\end{equation}
From (\ref{2thirteen}), one can immediately verify that the metric derived above has the correct Newtonian limit---the
potential reduces to the classic Hill potential. This solution for the even parity solutions was first
discovered by Moeckel~\cite{Mo}.
The entire metric may now be written out, defining $P \equiv P_2(\cos\theta) =
3\cos^2\theta - 1$:
\begin{equation}
\label{2fourteen}
\begin{array}{l}
\mathflushleft{ds^2 = \Bigl[-\bigl(1-{2M \over r}\bigr)+{m \over R^3}P(r-2M )^2\Bigr]dt^2}\\
\mathstrutter{15pt}\mathcenter{+ \Bigl[\bigl(1-{2M \over r}\bigr)^{-1}+{m \over R^3}Pr^2\Bigr]dr^2}\\
\mathstrutter{15pt}\mathflushright{+ \Bigl[r^2 + {m \over R^3}Pr^2(r^2-2M^2)\Bigr]d\Omega^2 \, .}
\end{array}
\end{equation}
We note that the perturbed spacetime (\ref{2fourteen}) furnishes another example supporting the
conjecture of Bekenstein ~\cite{Bek} that the horizon area of a near-equilibrium
black hole is an adiabatic invariant (see Appendix).
To compare with the hybrid-PN treatment, we transform to harmonic coordinates
by replacing the areal radial coordinate, $r$, with the harmonic radius, $r_h = r - M$.
Dropping the subscript, we obtain in harmonic coordinates
\begin{equation}
\label{2fifteen}
\begin{array}{l}
\mathflushleft{ds^2 \ = \displaystyle -{r-M \over r+M }dt^2 + {r+M \over r-M }dr^2 + (r+M)^2d\Omega^2}\\
\maththreefourths{\mathstrutter{15pt} \displaystyle + {m \over R^3}P\bigl[(r-M)^2dt^2 + (r+M)^2dr^2} \\
\mathflushright{\displaystyle + (r+M )^2(r^2+2rM -M^2)d\Omega^2\bigr] \, .}\\
\end{array}
\end{equation}
Obtaining the geodesic equations in this spacetime
up to first order in the tidal expansion parameter ${m r^2 \over R^3}$,
we find the equations of motion for a test particle near the black hole to be
\begin{equation}
\label{2nineteen}
\begin{array}{l}
\displaystyle
\mathflushleft{\ddot r = {3M \dot r^2 \over r^2-M^2}-{M (r-M ) \over (r+M )^3} + (r-M )\Omega^2}\\
\mathonefourth{\mathstrutter{15pt} + {m \over R^3} P \Big[-3r\dot r^2+(r+2M ){(r-M )^2 \over (r+M )^2}}\\
\mathcenter{\mathstrutter{15pt} + (r-M)(r^2+4rM+M^2)\Omega^2\Big] }\\
\mathflushright{\mathstrutter{15pt} - {m \over R^3} 2(r^2-M^2){dP \over d\theta} \dot\theta\dot r}\\
\\
\mathflushleft{\ddot\theta = {4M-2r \over r^2-M^2}\dot\theta\dot r + \cos\theta\sin\theta \dot\phi^2}\\
\mathonefourth{\mathstrutter{15pt} + {m \over R^3} P [-2(M+2r)\dot\theta\dot r]}\\
\mathcenter{\mathstrutter{15pt} + {m \over R^3} {dP \over d\theta} {1 \over 2} \Big[-(3r^2+2rM-3M^2)\dot\theta^2}\\
\mathflushright{\mathstrutter{15pt} + (r^2+2rM-M^2)\sin^2\theta \ \dot\phi^2 + {(r-M)^2 \over (r+M)^2} + \dot r^2\Big]}\\
\\
\mathflushleft{\ddot\phi \ =\ {4M-2r \over r^2-M^2} \dot\phi\dot r - 2 \cot\theta \ \dot\theta\dot\phi - 2{m \over R^3} P(M+2r)\dot\phi\dot r}\\
\mathflushright{\mathstrutter{15pt} -2{m \over R^3} {dP \over d\theta} (r^2+rM-M^2)\dot\theta\dot\phi \, ,}\\
\end{array}
\end{equation}
where dot denotes differentiation with respect to Schwarzschild coordinate time,
and $\Omega^2 = \dot\theta^2 + \sin^2\theta\dot\phi^2$.
To avoid coordinate singularities, it is convenient to integrate (\ref{2nineteen}) in
Cartesian coordinates. To facilitate an otherwise tedious transformation, we first
rewrite the equations of motion in 3-vector form.
We make the identification
\begin{equation}
\label{defacc}
\begin{array}{lll}
{\bf \ddot r} = & &( \ddot r-r\dot\phi^2\sin^2\theta-r\dot\theta^2){\partial\over\partial r}\\
\mathstrutter{15pt} &+ &( \ddot\theta+(2 \dot r\dot\theta-r\dot\phi^2\sin\theta \cos\theta)/r ){\partial \over \partial\theta}\\
\mathstrutter{15pt} &+ &( \ddot\phi +(2 \dot r\dot\phi \sin\theta+2r\dot\theta\dot\phi \cos\theta)/(r\sin\theta) ){\partial\over\partial\phi} \, ,\\
\end{array}
\end{equation}
a vector which we have constructed to be the same as the Newtonian 3-acceleration in spherical
coordinates and we know transforms to $(\ddot x,\ddot y,\ddot z)$ in Cartesian coordinates. Identifying
the velocity 3-vector ${\bf v} = (\dot r,\dot\theta,\dot\phi)$ in the spherical
coordinate basis and setting ${\bf r} = r{\partial\over\partial r}$ allows us to write
(\ref{2nineteen}) in the compact form 3-vector form
\begin{equation}
\label{rvector}
\begin{array}{l}
\displaystyle
\mathflushleft{{\bf \ddot r} = -{M \over r^2} [A{\bf e_r} + B{\bf v}]} \\
\displaystyle
\mathflushright{\mathstrutter{15pt} + {m \over R^3 } \left[\alpha \left({ 3{\bf r}\cdot {\bf R} \over R^2 }{\bf R} - {\bf r}\right) + \beta {\bf v} + \gamma {\bf e_r} \right] \, ,}\\
\end{array}
\end{equation}
where $A$ and $B$ are again given by Eqs.~(\ref{twelve}) and ~(\ref{thirteen}).
The quantities $\alpha$, $\beta$, and $\gamma$ are given by
\begin{equation}
\label{2twentyone}
\begin{array}{l}
\displaystyle
\mathflushleft{\alpha = (r^2+2rM-M^2){v^2-\dot r^2 \over r^2} + \dot r^2 + {(r-M)^2 \over (r+M)^2}} \\
\\
\displaystyle
\mathflushleft{\beta = -2(3\cos^2\Theta-1)(M+2r)\dot r} \\
\displaystyle
\mathflushright{\mathstrutter{15pt} + 12\cos\Theta\sin\Theta(r^2+rM-M^2)\dot\Theta}\\
\\
\displaystyle
\mathflushleft{\gamma = (3\cos^2\Theta-1)\Biggl[2M\bigl(\dot r^2+{(r-M)^2 \over (r+M)^2} \bigr)}\\
\displaystyle
\maththreefourths{\mathstrutter{15pt} + M(r^2-2rM-M^2) {v^2-\dot r^2 \over r^2} \Biggr]}\\
\displaystyle
\mathflushright{\mathstrutter{15pt} - 12rM \dot r\dot\Theta\cos\Theta\sin\Theta \, ,}\\
\end{array}
\end{equation}
where $\Theta$ is the angle between $\bf r$ and ${\bf R}$. It is now straightforward
to write out the Cartesian components of Eq.~(\ref{rvector}).
To incorporate the (slow) inspiral of the
companion in the context of our quasi-static approximation, we treat ${\bf R} = {\bf R(t)}$
in Eq. (\ref{rvector}) as a
parameter, which is slowly evolved in accord with
Eqs. (\ref{fifteen}), (\ref{sixteen}) and (\ref{seventeen}).
The leading terms independent of the companion $m$ are identical in
both (\ref{rvector}) and (\ref{six}), a result
which is consistent with the adoption of hybrid equations
in the PN analysis. The tidal term in
(\ref{rvector}) reduces to the lowest-order Newtonian expression used in
(\ref{six}) in the weak-field, slow-velocity limit where
$ v \ll 1$, $ M/r \ll 1$. Since the test particles orbit close to the hole, the
relativistic tidal expression will indeed cause departures from the motion
predicted by the Newtonian tidal term.
\subsection{Numerical Implementation}
A spherical swarm identical to the one evolved with the hybrid-PN equations
of motion was evolved with Eq.(\ref{rvector}). Because collapse occurs later
in this simulation, the swarm was evolved until $R/M=26.30$. The results are summarized
in Figures 3 -- 5.
In Figure 3, the mean cluster radius is computed once again and compared to
the hybrid-PN result. The qualitative nature of the
evolution is the same, and the existence of a crushing instability is evident.
The main effect of the new relativistic terms is to delay the
collapse until the perturber gets somewhat closer to the hole.
In particular, of 20,000 particles, 16,324 fell into the hole (81.6 \%) by the time
the companion reached $R/M = 26.30$, a slightly smaller percentage than in the hybrid-PN simulation.
Once again, the compact cluster is observed to be stable in isolation, but
driven to collapse by the presence of a sufficiently strong tidal perturbation.
\begin{figure}
\epsfxsize=3in
\begin{center}
\leavevmode
\epsffile{fig3a.ps}
\end{center}
\epsfxsize=3in
\begin{center}
\leavevmode
\epsffile{fig3b.ps}
\end{center}
\caption{Evolution of the cluster in the Schwarzschild perturbation treatment.
The mean radius of the 20,000 test-particles in the initially spherical swarm
is plotted as a function of the tidal companion radius in the top figure and
as a function of coordinate time in the bottom figure. In the top figure we
compare the results found for our two treatments; the behavior is qualitatively
similar, but the crushing effect occurs slightly later in the fully relativistic,
Schwarzschild perturbation treatment.}
\end{figure}
\onecolumn
\begin{figure}[p]
\begin{center}
\leavevmode
\hbox{\vbox{
\hbox{\hbox to 3in{\hfil \Huge \bf A \hfil} \hbox to 1in{} \hbox to 3in{\hfil \Huge \bf B \hfil}}
\hbox{\vrule width 0pt height 5pt}
\hbox{\fbox{\epsfxsize=3in\epsffile{fig4a.ps}
\hbox to 1in{}
\epsfxsize=3in\epsffile{fig4b.ps}}
\hskip-7in
\vbox to 2.25in{\vskip5pt \hbox to 7in{\hfil \LARGE $t/M = 0$ \hfil} \vfil}
}
\hbox{\fbox{\epsfxsize=3in\epsffile{fig4c.ps}
\hbox to 1in{}
\epsfxsize=3in\epsffile{fig4d.ps}}
\hskip-7in
\vbox to 2.25in{\vskip5pt \hbox to 7in{\hfil \LARGE $t/M = 1.18\times10^5$ \hfil} \vfil}
}
\hbox{\fbox{\epsfxsize=3in\epsffile{fig4e.ps}
\hbox to 1in{}
\epsfxsize=3in\epsffile{fig4f.ps}}
\hskip-7in
\vbox to 2.25in{\vskip5pt \hbox to 7in{\hfil \LARGE $t/M = 1.22\times10^5$ \hfil} \vfil}
}
\hbox{\vrule width 0pt height 5pt}
}}
\end{center}
\caption{Snapshots of the cluster swarm at selected times during the binary inspiral.
In $A$ we show the view looking down along the $z$-axis, which is perpendicular to
the orbital plane of the companion; in $B$ we show the view looking along the $x$-axis,
which lies in the orbital plane of the companion. For the
top frame the companion is at $R/M = 60$; for the middle frame the companion
is at $R/M = 30.7$; for the bottom frame the companion is at $R/M = 26.30$.}
\end{figure}
\twocolumn
\begin{figure}
\begin{center}
\leavevmode
\hbox to 246pt{\hskip40pt
\hbox{\vbox{
\hbox{\fbox{\epsfxsize=2in\epsffile{fig5a.ps}}
\hskip-2in
\vbox to 1.5in{\vskip5pt \hbox{\LARGE a} \vfil}}
\hbox{\fbox{\epsfxsize=2in\epsffile{fig5b.ps}}
\hskip-2in
\vbox to 1.5in{\vskip5pt \hbox{\LARGE b} \vfil}}
\hbox{\fbox{\epsfxsize=2in\epsffile{fig5c.ps}}
\hskip-2in
\vbox to 1.5in{\vskip5pt \hbox{\LARGE c} \vfil}}
\hbox{\vrule width 0pt height 5pt}
}}
\hfil}
\end{center}
\caption{ Test-particle trajectories about the black hole during the
inspiral of the binary companion from $R/M=60$ to $R/M=26.30$.
Frame (a) shows the orbit of a particle moving in the orbital plane
of the companion; it is not captured by the time the integrations terminate.
Frame (b) shows the capture of a particle initially orbiting in
a plane perpendicular to the companion plane. Frame (c)
shows the precession and capture of a particle initially orbiting at
an angle of $45^o$ to the companion plane.}
\end{figure}
As the geodesic equations in the perturbative relativistic treatment are
derived from a self-consistent Lagrangian, they satisfy strict conservation laws
even in the presence of a companion, assuming it is stationary. These conservation
laws provide a means of testing our code and particle integration scheme.
For example, if the companion is fixed at an arbitrary position on the
z-axis, the perturbed Schwarzschild spacetime (\ref{2fifteen}) admits two Killing vectors,
$\partial\over\partial t $ and $\partial\over\partial \phi$, yielding
conservation of particle energy $p_t$ and angular momentum $p_{\phi}$,
even for large tidal fields.
Given that we linearize the tidal field and retain only the lowest order terms in $m$ in our
equations of motion (\ref{rvector}),
energy and angular momentum conservation are no longer exact. However, we have
tested our code and have shown that it reliably obeys
these conservation laws to the required order.
The evolution of the swarm is depicted in Figure 4, where snapshots of the
cluster are shown at three different times from two different viewing angles.
By the end of the simulation,
the original spherical swarm is reduced to a sparse cylindrical band of particles whose
axis is perpendicular to the orbital plane of the companion.
Apparently, particle orbits at small inclination angles to this plane are more stable
than those which are perpendicular, a result already noted in
~\cite{Us}. This feature is evident in Figure 5, which plots the trajectories of three
representative particles. The simulation confirms that gravitational collapse can be induced
in a collisionless cluster by the tidal field of a binary companion.
|
\section{Introduction}
Classical continuum Heisenberg ferromagnet models (CCHFM)
exhibit a
rich variety of nonlinear behaviour. In particular, over the past two
decades, several integrable nonlinear ferromagnetic models have
been idenfied [1-3]. So, some number integrable spin systems
in 2+1 dimensions are found [4-8]. The
underlying nonlinear spin excitations in such spin systems represented
by solitons, domain
walls, vortecies,
lumps and dromions [4-11]. Theirs study are of considerable intrinsic interest,
especially from the points of view of both mathematics and physics.
Integrable spin systems in 2+1 dimensions as their (1+1)-dimensional
counterparts, display
fascinating geometrical aspects: they are gauge and L-equivalent to the
nonlinear
Schr${\ddot o}$dinger - type equations (the Davey-Stewartson equation,
the Zakharov equations (ZE), the Strachan equation and so on) [5, 9-10, 12].
Generally speaking, between spin systems and the differential geometry take
places the deep connection [12, 13-17, 20]. An important feature of two dimensional
spin systems is the existence the topological invariant
$$
Q = \frac{1}{4\pi} \int\int dxdy {\bf S}\dot ({\bf S}_{x} \wedge {\bf S}_{y}) \eqno (1)
$$
such that the solutions of them are classified by the topological charge (1).
In (9) ${\bf S} = (S_{1}, S_{2}, S_{3})$ is the three-dimensional unit vector
(the spin vector). The class of exact solutions of (2+1)-dimensional spin
systems is a very rich one. The solitons, vortices, dromions, lumps
are among them. These solutions have the important physical significance.
So, for example, vortecies play an active role in the dynamics and
termodynamics of quasi-two-dimensional magnets [18-20].
The present paper is devoted to the study of the
following CCHFM in two
space and one time dimensions
$$
{\bf S}_t + {\bf S}\wedge \{(b+1) {\bf S}_{\xi \xi} -b{\bf S}_{\eta \eta}\} +
bu_{\eta} {\bf S}_{\eta} + (b+1)u_{\xi}{\bf S}_{\xi} = 0 \eqno(2a)
$$
$$
u_{\xi \eta} = {\bf S}({\bf S}_{\xi}\wedge {\bf S}_{\eta}) \eqno(2b)
$$
with
$$
\xi = \frac{x}{2} + \frac{a+1}{\alpha}y, \quad \eta = -\frac{x}{2} -
\frac{a}{\alpha}y. \eqno(3)
$$
where $a, b $ are real constants, $\alpha^{2} = \pm 1$.
It is the Myrzakulov XX (M-XX) equation [5].
Hereafter, for convenience, we use
the conditional notations, e.g. equation (2) we denote by
the M-XX equation.
Equation (2) is integrable. We will distinguish the two integrable cases:
the M-XXA equation as
$\alpha^{2} = 1$ and the M-XXB equation as $\alpha^{2} = -1$. Also,
equation (2) contains several integrable cases: \\
(i) $b = 0$, yields the M-VIII equation
$$
{\bf S}_t + {\bf S}\wedge {\bf S}_{\xi \xi} + w{\bf S}_{\xi} = 0 \eqno(4a)
$$
$$
w_{ \eta} = {\bf S}({\bf S}_{\xi}\wedge {\bf S}_{\eta}) \eqno(4b)
$$
where $w=u_{\xi}$.
(ii) $a = b = - \frac{1}{2}$, yields the celebrated Ishimori equation [4]
$$
{\bf S}_t + \frac{1}{2}{\bf S}\wedge \{ {\bf S}_{\xi \xi} +
{\bf S}_{\eta \eta}\} -\frac{1}{2}u_{\eta} {\bf S}_{\eta} + \frac{1}{2}u_{\xi}{\bf S}_{\xi} = 0 \eqno(5a)
$$
$$
u_{\xi \eta} = {\bf S}({\bf S}_{\xi}\wedge {\bf S}_{\eta}) \eqno(5b)
$$
where $\xi = \frac{1}{2}(x + \frac{1}{\alpha}y), \quad \eta =
-\frac{1}{2}(x -\frac{1}{\alpha}y).$ The Ishimori equation (5) is the first
integrable spin system in plan, which can be solved by the inverse scattering
method (IST). This equation were studied by the different authors from
variety points of view (see, e.g. [4,6-7,9,11,21]).
(iii) Let now we put $b = 0, \eta =t$, then equation (2) reduces to
the (1+1)-dimensional M-XXXIV equation
$$
{\bf S}_t + {\bf S}\wedge {\bf S}_{\xi \xi} + w{\bf S}_{\xi} = 0 \eqno(6a)
$$
$$
w_{ t} + \frac{1}{2} ({\bf S}^{2}_{\xi})_{\xi} = 0 \eqno(6b)
$$
This integrable equation was introduced in [5] to describe nonlinear dynamics of
compressible magnets.
Equation (2) is the (2+1)-dimensional
integrable generalization of the (1+1)-dimensional CCHFM or the isotropic
Landau-Lifshitz equation (LLE)
$$
{\bf S}_t = {\bf S} \wedge {\bf S}_{xx} \eqno(7)
$$
and in 1+1 dimensions reduces to it. Here, it should be mentioned that
the M-XX equation (2) is not the only integrable generalization of
the LLE (7) in 2+1 dimensions. There exist several another integrable
generalizations, e.g the following one,
$$
{\bf S}_t = ( {\bf S} \wedge {\bf S}_{y} + u {\bf S})_x \eqno(8a)
$$
$$
u_{x} = - {\bf S}({\bf S}_x \wedge {\bf S}_y) \eqno(8b)
$$
This equation, which is known as the Myrzakulov I (M-I) equation,
is again completely integrable [5,10,12].
As integrable, equation (2) can be solved by the IST method.
The applicability of the IST method to the M-XX
equation (2) is based on its equivalence to the compatibility condition
of the following linear equations [5]
$$
\Phi_{Z^{+}} = S\Phi_{Z^{-}} \eqno(9a)
$$
$$
\Phi_{t} = 2i[S+(2b+1)I]\Phi_{Z^{-}Z^{-}} + W\Phi_{Z^{-}} \eqno(9b)
$$
where $Z^{\pm} = \xi \pm \eta$ and
$$ W = 2i\{(2b+1)(F^{+} + F^{-} S) +(F^{+}S + F^{-}) +
(2b+1)SS_{Z^{-}}+\frac{1}{2}S_{Z^{-}} + \frac{1}{2} SS_{Z^{+}} \}, \quad
$$
$$
S= \pmatrix{
S_3 & rS^- \cr
rS^+ & -S_3
},\quad S^{\pm}=S_{1}\pm iS_{2} \quad S^2 = EI,\quad E = \pm 1,
\quad r^{2}=\pm 1,
$$
$$
F^{+} = 2iu_{Z^{-}}, \quad F^{-}=2iu_{Z^{+}}.
$$
In fact, from the condition $\Phi_{Z^{+}t} = \Phi_{tZ^{+}}$, we get the equation
$$
iS_t + \frac{1}{2}[S,(b+1) S_{\xi \xi} -bS_{\eta \eta}] +
ibu_{\eta} S_{\eta} + i(b+1)u_{\xi}S_{\xi} = 0 \eqno(10a)
$$
$$
u_{\xi \eta} = \frac{1}{4i}tr(S[S_{\xi},S_{\eta}]) \eqno(10b)
$$
which is the matrix form of equation (2).
\section{Bilinearization}
It could be of interest to study the equation (2) by the IST method. But
to look for the some special solutions,
it is convenient use the other may be more practical method - the Hirota
bilinear method. Remaining the use of the IST method in future, in this
paper, we work with the Hirota method. To this purpose, we construct the
bilinear form of (2) as $E=r=1$. Let us we now introduce the following
transformation for the components of spin vector ${\bf S}$
and for the derivatives of scalar potential $u$
$$
S^{+} = \frac{2\bar f g}{\bar f f + \bar g g}, \quad
S_3 = \frac{\bar f f - \bar g g}{\bar f f + \bar g g}, \eqno (11a)
$$
$$
u_{\xi} =- 2i\frac{D_{\xi}(\bar f\circ f +
\bar g \circ g)}{\bar f f +
\bar g g}, \quad
u_{\eta}= 2i \frac{D_{\eta}(\bar f \circ f +
\bar g \circ g)}{\bar f f + \bar g g} \eqno (11b)
$$
Here the Hirota operators $D_{x}, D_{y}$ and $D_{t}$ are defined by
$$
D^{l}_{\xi}D^{m}_{\eta}D^{n}_{t} f(\xi, \eta, t)\circ g(\xi, \eta, t) =
(\partial_{\xi}-\partial_{\xi^{\prime}})^{l}
(\partial_{\eta}-\partial_{\eta^{\prime}})^{m}
(\partial_{t}-\partial_{t^{\prime}})^{n} f(\xi, \eta, t)\circ g(\xi^{\prime},
\eta^{\prime}, t^{\prime})\mid_{\xi=\xi^{\prime}, \eta = \eta^{\prime},
t=t^{\prime}} . \eqno(12)
$$
Substituting the formulae (11) into the M-XX equation (2), we obtain
the bilinear equations
$$
[iD_{t} - (b+1)D_{\xi}^{2} +bD_{\eta}^{2}](\bar f \circ g) = 0 \eqno (13a)
$$
$$
[iD_{t} - (b+1)D_{\xi}^{2} +bD_{\eta}^{2}](\bar f \circ f - \bar g \circ g) = 0 \eqno (13b)
$$
$$
\{D_{\xi} D_{\eta}+D_{\eta} D_{\xi}\} (\bar f f + \bar g g) \circ
(\bar f f + \bar g g) = 0 \eqno (13c)
$$
Note that equation (13c) coincide with the compatibility condition $ u_{\xi\eta} = u_{\eta\xi} $.
Equations (13) is the desired Hirota bilinear form of equation (2).
\section{Solutions}
Now we can construct some special solutions of equation (2). As examples,
we find simplest soliton, domain wall and vortex solutions.
\subsection{Soliton solution}
\begin{center}
{\bf FIND THE SOLITON SOLUTIONS.}
\end{center}
\subsection{Domain wall solution}
In order to obtain a domain wall solutions,
we make the choice
$$
f=1. \eqno (14)
$$
Then, equations (13a,b) reduce to
$$
ig_{t}+(b+1)g_{\xi\xi} - b g_{\eta\eta} = 0 \eqno(15a)
$$
$$
(b+1)\bar g_{\xi}g_{\xi} - b\bar g_{\eta}g_{\eta} = 0 \eqno(15b)
$$
Let us we consider the case, when $\alpha^{2} = -1$, i.e. the M-XXB equation.
In this case, equation (13c) is identically satisfied by any analytical
function $g = g(\xi, \eta, t)$. For example, the simplest non-trivial solution
of equation (2) is
$$
g = \exp \chi_{1} \eqno (16)
$$
where
$$
\chi_{1} = m_{1}\xi + n_{1}\eta + i [(b+1)m^{2}-bn^{2}]t + \chi_{10} =
\chi_{1R}+i\chi_{1I} \eqno (17)
$$
Thus, the spin components and the potential field are given by
$$
S^{+} = e^{i\chi_{1I}} sech \chi_{1R},\quad
S_{3} = - th\chi_{1R},\quad u = 2\ln (1+e^{2\chi_{1R}}). \eqno (18)
$$
\subsection{ Vortex solution}.
To construct vortex solution, we use the equation (13) and assume that its
solution has the form
$$
f=f(\xi, t), \quad g = g(\xi, t) \eqno(19)
$$
Then the condition (13c) is satisfied automatically. At the same time,
equations (13a,b) are satisfy if
$$
if_{t}+ (b+1) f_{\xi\xi} = 0 \quad ig_{t} + (b+1) g_{\xi \xi} = 0 \eqno(20)
$$
Hence, we obtain the following multi-vortex solutions of the M-XXB equation
(2)
$$
g_{N} = \sum_{j=0}^{N}\sum_{m+2n=j}\frac{a_{j}}{m!n!}(\frac{2}{b+1})^{\frac{m}{2}}\xi^{m}(2it)^{n} \eqno(21a)
$$
$$
f_{N} = \sum_{j=0}^{N-1}\sum_{m+2n=j}\frac{b_{j}}{m!n!}(\frac{2}{b+1})^{\frac{m}{2}}\xi^{m}(2it)^{n} \eqno(21b)
$$
where $a_{j}$ and $b_{j}$ are arbitrary complex constants, $m,n$ are the
non-negative integer numbers. In particular, the 1-vortex solution isgiven by
$$
f=b_{0}, \quad g=a^{\prime}_{1}\xi + a_{0} \eqno (22)
$$
where $a^{\prime}_{1}=a_{1}(\frac{2}{b+1})^{\frac{1}{2}}$. The corresponding solution
of equation (2) is given by
$$
S^{+} = \frac{2b_{0}(a^{\prime}_{1}\xi + a_{0})}{\mid b_{0}\mid^{2} +
\mid a^{\prime}_{1}\xi + a_{0}\mid^{2}}
\eqno (23a)
$$
$$
S_{3} = \frac{
\mid b_{0}\mid^{2} - \mid a^{\prime}_{1}\xi + a_{0}\mid^{2}}
{\mid b_{0}\mid^{2} + \mid a^{\prime}_{1}\xi + a_{0}\mid^{2}}
\eqno (23b)
$$
$$
u = 2\ln (\mid b_{0}\mid^{2} + \mid a^{\prime}_{1}\xi + a_{0}\mid^{2})
\eqno (23c)
$$
So, the 1-vortex solution is static. To find the dynamic solution,
we consider the 2-vortex solution, which has the form
$$
f=b^{\prime}_{0}\xi + b_{0}, \quad g=\frac{a_{2}}{b+1}\xi^{2} +
\frac{a_{2}}{2} 2it+
a^{\prime}_{1}\xi + a_{0}, \quad b^{\prime}_{1}=b_{1}(\frac{2}{b+1})^{\frac{1}{2}}
\eqno (24)
$$
The interesting question is the dynamics of these vortices. Let
us rewrite the solution (21) in the following factorized form
$$
f(\xi, t) = b_{0} \prod^{N}_{j=1}[\xi - p_{j}(t)] \eqno(25a)
$$
$$
g(\xi, t) = a_{0} \prod^{N}_{j=1}[\xi - q_{j}(t)] \eqno(25b)
$$
where $p_{j}$ and $q_{j}$ denote the positions of the zeros for $f$ and $g$,
and $a_{0}, b_{0}$ are constants. Substituting (25) into (20), we get the
evolutions of $p_{j}$ and $q_{j}$ as
$$
p_{jt} = -i(b+1)\sum^{N}_{k\not= j}\frac{1}{p_{j}-p_{k}} \eqno(26a)
$$
$$
q_{jt} = -i(b+1)\sum^{N}_{k\not= j}\frac{1}{q_{j}-q_{k}} \eqno(26b)
$$
where $j,k=1,2,...,N$. Hence, we get the Calogero-Moser type system
$$
p_{jtt} = 2(b+1)^{2}\sum^{N}_{k\not= j}\frac{1}{(p_{j}-p_{k})^{3}} \eqno(27a)
$$
$$
q_{jtt} = 2(b+1)^{2}\sum^{N}_{k\not= j}\frac{1}{(q_{j}-q_{k})^{3}} \eqno(27b)
$$
with the following Hamiltonian
$$
H=\frac{1}{2} \sum (p_{jt}^{2}+q^{2}_{jt}) + (b+1)^{2}\sum [ (p_{j}-p_{k})^{-2}
+(q_{j}-q_{k})^{-2}]. \eqno(28)
$$
\subsection{ Dromion solution}
In this subsection we would like get the dromion [23] solution of equation (2),
but please
\begin{center}
{\bf FIND THE DROMION SOLUTIONS.}
\end{center}
\section{Gauge equivalent equation}
Finally, let us we present the gauge equivalent counterpart of equation (2).
It has the form
$$
iq_t+ (1 + b)q_{\xi \xi } - b q_{\eta \eta } + vq = 0 \eqno(29a)
$$
$$
ip_t - (1 + b)p_{\xi \xi } + b p_{\eta \eta } - vp = 0 \eqno(29b)
$$
$$
v_{\xi \eta } = -2\{(1+ b) (pq)_{\xi \xi} - b(pq)_{\eta \eta}\} \eqno(29c)
$$
where $p,q$ are some complex functions. Equation (29) is related with the
Zakharov equations [22]. To prove gauge equivalence between equations
(2) and (29), let us perform the
gauge transformation $\Psi = g \Phi$, where the function $\Phi$ is
the solution of equations (9) and $g$ is a 2x2 matrix such that
$$
S = g^{-1}\sigma_{3}g \eqno (30)
$$
and
$$
g_{Z^{+}}g^{-1} - \sigma_{3} g_{Z^{-}}g^{-1}
=
\left ( \begin{array}{cc}
0 & q \\
p & 0
\end{array} \right) \eqno(31)
$$
Under this transformation the function $\Psi$ satisfies the following
set of linear equations
$$
\Psi_{Z^{+}} = \sigma_{3} \Psi_{Z^{-}} + B_{0}\Psi \eqno(32a)
$$
$$
\Psi_{t} = 4i C_{2} \Psi_{Z^{-}Z^{-}} + 2 C_{1} \Psi_{Z^{-}} + C_{0}\Psi. \eqno(32b)
$$
where $B_{0},\quad C_{j}$ are given by
$$
B_{0}= \pmatrix{
0 & q \cr
p & 0
}, \quad
C_{2}= \pmatrix{
b+1 & 0 \cr
0 & b
},\quad
C_{1}= \pmatrix{
0 & iq \cr
ip & 0
},\quad
C_{0}= \pmatrix{
c_{11} & c_{12} \cr
c_{21} & c_{22}
}
$$
$$
c_{12}=i[(4b+3)q_{Z^{-}}+q_{Z^{+}}] \quad
c_{21}=-i[(4b+1)p_{Z^{-}}+ p_{Z^{+}}]
$$
and $v=i(c_{22}-c_{11})$. Here $c_{jj}$ are the solution of the following
equations
$$
c_{11Z^{-}}- c_{11Z^{+}} = i[(4b+3)(pq)_{Z^{-}} + (pq)_{Z^{+}}]
$$
$$
c_{22Z^{-}}+ c_{22Z^{+}} = i[(4b+1)(pq)_{Z^{-}} + (pq)_{Z^{+}}].
$$
The compatibility condition of equations (32) gives the equation (29). Therefore
the M-XX equation (2) and the equation (29) are gauge equivalent to each other.
Now let us proceed to the M-XX equation (10). It is not difficult to
check that if $g$ obeys equations (31) then the $S$ in the form (30) obeys
the M-XX equation (10) with
$$
u = -2i\alpha \ln \det g \eqno(33)
$$
\section{Conclusion}
To conclude, in this paper we have found some exact solutions, namely
domain wall and vortex solutions of the (2+1)-dimensional CCHFM - the
M-XX equation. We have shown that the dynamics of vortices are governed by the
Calogero-Moser type systems. Also we have presented the gauge equivalent counterpart of this
equation.
\section{Particular open problems}
Finally, we also would like to pose the following particular problems: \\
{\bf Problem N1:} Find solutions of the M-XX equation by
the $\bar \partial$-dressing method. \\
{\bf Problem N2:} Find solutions of the M-XX equation using
the nonlocal Riemann-Hilbert problems method. \\
{\bf Problem N3:} Find solutions of the M-XX equation by
means of the Darboux transformation as in [21]. \\
{\bf Problem N4:} Find the other solutions of the M-XX equation
(solitons, dromions and so on) by the Hirota bilinear method.\\
If you have some results in these directions, please, inform me by
E-mail: <EMAIL>. We are ready to interaction.
\section{Acknowledgments}
The author would like to thank Prof. M.Lakshmanan for hospitality
during his visit to Bharathidasan University and for stimulating
discussions.
He is grateful for helpful conversations with
A.Kundu, Radha Balakrishnan, M.Daniel, R.Radha and R.Amuda.
Also he would like to thank Dr.
Radha Balakrishnan for hospitality during his visit to the Institute of
Mathematical Sciences.
|
\section{Introduction}
The compact non-thermal radio source Sgr~A* has long been recognized as
a massive black hole candidate. Recent results from
stellar proper motion studies indicate that there is a
dark mass of $\sim 2.6 \times 10^6 M_{\sun}$ enclosed within 0.01 pc
(Genzel {\it et al.\ } \markcite{genze97} 1997, Ghez {\it et al.\ } \markcite{ghez98} 1998).
Very long baseline interferometry studies at millimeter wavelengths
have shown that the intrinsic radio source coincident with the dark mass
has a size that is less than
1 AU and a brightness temperature greater than $10^9$ K (Rogers {\it et al.\ }
\markcite{roger94} 1994,
Bower \& Backer \markcite{bower98} 1998, Lo {\it et al.\ } \markcite{lo98} 1998). Together these points are compelling
evidence that Sgr~A* is a cyclo-synchrotron emitting region surrounding
a massive black hole. Nevertheless, specific details of the excitation of high
energy electrons, their distribution and the accretion of infalling
matter onto Sgr~A* are unknown (e.g., Falcke, Mannheim \& Biermann \markcite{falck93} 1993,
Melia \markcite{melia94} 1994, Narayan {\it et al.\ } \markcite{naray98} 1998, Mahadevan \markcite{mahad98} 1998).
Linear polarization stands as one of the few observables of Sgr~A*
not extensively investigated observationally or theoretically. However,
we expect linear polarization to arise from the cyclo-synchrotron
radiation that is responsible for the radio to millimeter wavelength
spectrum. A homogeneous, optically-thin, synchrotron
source with a uniform magnetic
field has a fractional polarization of 70\%. Measured fractional
polarizations in AGN are typically a few percent at
wavelengths shorter than 6 cm where the compact cores dominate
the highly-polarized radio lobes in
the total flux (e.g., Aller, Aller
\& Hughes \markcite{aller92} 1992).
However, polarization VLBI images sometimes show regions
of significantly enhanced polarization
(Brown, Roberts \& Wardle \markcite{brown94} 1994).
The polarization of Sgr~A* may prove to be as important a diagnostic of
models for the radio to millimeter spectrum as it has been for AGN.
Detection of linear polarization in AGN has firmly established
synchrotron emission as the radiation mechanism.
Comparison of the evolution
of linear polarization to the evolution of total intensity has provided
a strong argument for the existence of shocks in the relativistic
jets of AGN (e.g., Hughes, Aller and Aller \markcite{hughe85} 1985).
Detection of similar
correlations in polarized and total intensity variations in Sgr~A* would be
convincing evidence for a jet. Other models may have unique
signatures for polarized intensity variations.
Sgr~A* is located in a region with strong magnetic fields and high
electron density. The image of Sgr~A* is significantly scatter-broadened
by intervening thermal plasma (e.g., Lo {\it et al.\ } \markcite{lo98} 1998), as are the images
of many masers in the Galactic Center region (Frail {\it et al.\ } \markcite{frail94} 1994).
Furthermore, nonthermal filaments in the Galactic
Center region show RMs which vary on the arcsecond scale and
are as large as 4000 ${\rm\ rad\ m^{-2}}$ (Yusef-Zadeh, Wardle \& Parastaran \markcite{yusef97} 1997).
Such large RMs can effectively depolarize a signal detected with a
large bandwidth.
In \S 2, we discuss the effect of large RMs on a polarized signal
and our Fourier transform technique for detecting large RMs.
In \S 3.1, we present VLA
continuum observations at 4.8 GHz.
In \S 3.2 and \S 3.3, we present VLA spectro-polarimetric observations
at 4.8
GHz and 8.4 GHz. These observations are sensitive to a wide-range of
RMs. In \S 4, we consider other effects of
interstellar matter on a polarized signal from Sgr~A*. And in
\S 5, we discuss the consequences of our upper limits for the
polarization on models for Sgr~A*. In a future paper, we will address
millimeter polarization observations of Sgr~A*.
\section{Searching for Large RMs}
In an ionized and magnetized region
right and left circularly polarized waves will have different
indices of refraction. This leads to a wavelength-dependent
delay between circular polarizations which
is equivalent to a rotation of the position angle $\chi$ of a
linearly polarized signal
\begin{equation}
\chi_F = {\rm RM}\ \lambda^2,
\end{equation}
where RM is the rotation measure.
This rotation of $\chi$ is equivalent to a rotation in the two-dimensional
Stokes
$Q$ and $U$ space.
A linearly polarized signal will be significantly depolarized in
an observing bandwidth $\Delta\nu$ if $\chi$ rotates by more than one
radian, or if the RM exceeds
\begin{equation}
{\rm RM_{max}}={1\over 2 } {1 \over \lambda^2}{\nu \over \Delta\nu}.
\label{eqn:rmmax}
\end{equation}
If the bandwidth $\Delta\nu$ is split into $n \times \delta\nu$ channels,
a search can be made for RMs larger than ${\rm RM_{max}}$.
When $\chi$ wraps through more than
one turn, $n\pi$ ambiguities make it impossible
through a linear least squares fit to detect RMs.
Fourier transforming
the complex visibility $P=Q + i U$ with respect to $\lambda^2$
searches for large RMs without loss of sensitivity.
The maximum RM detectable in
this scheme can be found by replacing $\Delta\nu$ with $\delta\nu$ in
Equation~\ref{eqn:rmmax}.
In addition to detecting RMs that exceed ${\rm RM_{max}}$, the technique
is sensitive to multiple RMs from the same object. A more detailed
analysis of this technique can be found in Killeen {\it et al.\ } \markcite{kille99} (1999).
A continuum observation with the VLA at 4.8 GHz with 50 MHz has
${\rm RM_{max}} \approx 10^4 {\rm\ rad\ m^{-2}}$.
Splitting the band into 256 channels increases
${\rm RM_{max}}$ by two orders of magnitude to
$3.5\times 10^6 {\rm\ rad\ m^{-2}}$.
The minimum fully-sampled RM detectable in a spectro-polarimetric data set,
${\rm RM_{min}}$, is
approximately equal to ${\rm RM_{max}}$ for a continuum data set with
the same total bandwidth.
The RM can be found to better accuracy than ${\rm RM_{min}}$. We estimate
the error to be
\begin{equation}
\sigma_{\rm RM} = { {\rm RM_{min} } \over {\rm SNR} } .
\end{equation}
SNR is the ratio of the peak amplitude in Fourier space to the off-peak
root-mean-square noise.
\section{Observations and Data Reduction}
\subsection{VLA Continuum Polarimetry at 4.8 GHz}
The VLA of the National Radio Astronomy Observatory\footnote{The
National Radio Astronomy Observatory is a facility
of the National Science Foundation operated under cooperative agreement
by Associated Universities, Inc.}
observed Sgr~A* on 10 and 18 April 1998 in the A array at 4.8 GHz
with a bandwidth of 50 MHz. Instrumental calibration was performed
with the compact sources 1741-038 and 1748-253. The right-left phase
difference was set with observations of 3C~286. Only baselines longer
than 100 $k\lambda$ were used for Sgr~A*. Several nearby calibrator
sources, GC 441, W56 and W109, were also observed (Backer \& Sramek \markcite{backe99} 1999).
All sources were self-calibrated and imaged in Stokes $I, Q$ and $U$.
We summarize the measured polarized and total intensities
of Sgr~A* and the calibrators in Table~\ref{tab:vla6cm}.
The rms noise in the Sgr~A* map is 74 $\ \mu {\rm Jy}$.
Consistency between the results on the two days indicates the
accuracy of the results. Polarization was reliably detected from
all sources but Sgr~A* and GC 441. The measured polarization
at the position of
Sgr~A* is 0.1\%. This value is equal to the average off source
fractional polarization in the map and is, therefore, an upper limit.
The maximum RM detectable with this bandwidth is $\sim 10^4 {\rm\ rad\ m^{-2}}$.
\subsection{VLA Spectro-polarimetry at 4.8 GHz}
The VLA observed Sgr~A* in the A array in a spectro-polarimetric mode
at 4.8 GHz on 27 November 1992. Observations were carried out in
8 consecutively-spaced frequency bands of 6.25 MHz each. Each band
was divided into 32 separate frequency channels. The bands
covered the frequency range from 4832 MHz to 4882 MHz. Five scans
of 2.5 minutes apiece on Sgr~A* were interleaved with six scans of
2.5 minutes apiece on NRAO~530 in each frequency band. Amplitude,
phase and polarization calibration were performed separately for each
band. Polarization calibration was performed with NRAO~530 alone
and with NRAO~530 and Sgr~A*, producing similar final results.
The right-left phase difference was set for each band with
an observation of 3C~286.
For each source, the spectral data were
time-averaged and exported from AIPS for further processing.
A bandpass correction was applied. The complex polarization was
then Fourier-transformed with respect to $\lambda^2$. Sampling
effects were removed through a one-dimensional CLEAN method.
The CLEAN method permits a better estimate of the RM peak
and of the noise level. The sampling sidelobes are readily
visible for 3C~286 and NRAO~530 in Figure~\ref{fig:fouramp6cm}.
Our tests with noise data and with synthetic signals indicate
that the CLEAN method does not generate false signals and
improves the accuracy of peak determination. Applying CLEAN
to the 4.8 GHz NRAO~530 data reduced the noise in the spectrum
from 1.6 mJy to 0.26 mJy.
The range of fully-sampled RM is $10^4 {\rm\ rad\ m^{-2}}$
to $3.5\times 10^6 {\rm\ rad\ m^{-2}}$. The Fourier amplitude for each
source is shown in Figure~\ref{fig:fouramp6cm} and the results
are summarized in Table~\ref{tab:rm6cm}. These images are without
bandpass correction and dirty-beam removal. We also calculate and
plot the Fourier transform for a distribution of Gaussian noise.
Strong peaks at low RM are apparent
for both 3C~286 and NRAO~530,
as expected. The measured values are
consistent with the known RMs of these sources: $1\pm 2 {\rm\ rad\ m^{-2}}$ for
3C~286 and $-63 \pm 5 {\rm\ rad\ m^{-2}}$ for NRAO~530 (Rusk \markcite{rusk88} 1988).
No strong peak is apparent for Sgr~A* at any RM. The maximum Fourier
amplitude for Sgr A* is $0.15\%$ at ${\rm RM}=2.1 \times 10^6 {\rm\ rad\ m^{-2}}$.
Imaging Sgr A* with and without a RM correction produced a peak polarization
of $0.2\%$. This is equal to the fractional polarization of
thermal ionized gas in the vicinity of Sgr A*, indicating that we are
limited by residual instrumental polarization.
\subsection{VLA Spectro-polarimetry at 8.4 GHz}
The VLA observed Sgr~A* in the A array in a spectro-polarimetric mode
at 8.4 GHz, also on 27 November 1992. Observations were carried out in
7 frequency bands of 6.25 MHz each. Each band
was divided into 32 separate frequency channels. Five bands
covered the frequency range from 8405 MHz to 8437 MHz. Two other
bands were centered at 8150 MHz and 8700 MHz. Five scans
of 2.5 minutes apiece on Sgr~A* were interleaved with six scans of
2.5 minutes apiece on NRAO~530 in each frequency band.
Amplitude,
phase and polarization calibration were performed separately for each
band. Polarization calibration was performed with NRAO~530 alone
and with NRAO~530 and Sgr~A* together, producing similar final results.
The right-left phase difference was set for each band with
an observation of 3C~286.
The sources W56, 1741-312, GC 441, W109
and 1748-253 were observed for two minutes in the three 6.25 MHz bands
centered at 8150 MHz, 8420 MHz and 8700 MHz.
The results for all sources were the same using all frequency bands or
only the inner 5 bands.
The same reduction steps were taken for the 8.4 GHz data as for the 4.8
GHz data.
The Fourier amplitudes for all sources are shown in
Figures~\ref{fig:fouramp4cm1} and \ref{fig:fouramp4cm2}.
These images are without
bandpass correction and dirty-beam removal.
The results are summarized in Table~\ref{tab:rm3cm}.
These data are sensitive to $3.5 \times 10^5 < | {\rm RM} | < 1.5 \times10^7 {\rm\ rad\ m^{-2}}$.
There are
strong detections of linear polarization in 3C~286 and NRAO~530 at
RMs consistent with zero.
Significant detections were also made for W56, W109, 1741-312 and 1748-253,
also at RMs consistent with zero. No polarization was detected in GC 441.
The errors in RM for these secondary calibrators are larger due to the
sparser frequency coverage and shorter observing time.
For Sgr A*, we detect a peak in the Fourier spectrum of 0.17\% at
${\rm RM}=24000 \pm 37000 {\rm\ rad\ m^{-2}}$.
Imaging Sgr A* with and without RM corrections,
we find a fractional polarization of 0.1\%. Off-source
fractional polarizations are typically 0.1\%, again implying that
we are limited by residual instrumental polarization.
We tested noise models to see if we could reproduce a weak signal at non-zero
RM. We used an input signal with ${\rm RM} = 0$ at 0.15\% of the peak
intensity of Sgr A* and noise that matched that of Sgr A*. This is
the model plotted in Figure~\ref{fig:fouramp4cm1}. The
measured RM peak wandered within the error range.
\section{Interstellar Propagation Effects}
The interstellar medium may depolarize a linearly
polarized radio wave in two ways:
significant rotation of the polarization position angle through the observing
bandwidth; and,
differential Faraday rotation along the many paths that contribute to
the scatter-broadened image of Sgr A*.
We have already addressed the first effect in \S 2 and found in
\S 3 that Sgr A* is not depolarized by RMs less than $1.5\times 10^7 {\rm\ rad\ m^{-2}}$.
We now consider the second effect.
The scattering region will depolarize the signal
if $\delta\chi_F\approx\pi$. For our observing wavelengths,
$\delta{\rm RM}=900 {\rm\ rad\ m^{-2}}$ and
$\delta{\rm RM}=2400 {\rm\ rad\ m^{-2}}$.
Over the scattering size of 50 mas at 4.8 GHz,
this corresponds to $\delta{\rm RM}/\delta{\theta}=18000 {\rm\ rad\ m^{-2}
arcsec^{-1}}$.
Observed variations in RM in the GC region are many orders of
magnitude below those necessary to depolarize Sgr A*.
Observations on the arcsecond to arcminute
scale of a nonthermal filament within 1 degree of Sgr A* find a maximum
$\delta{\rm RM}/\delta{\theta}=250 {\rm\ rad\ m^{-2}\ arcsec^{-1}}$
(Yusef-Zadeh, Wardle \& Parastaran \markcite{yusef97} 1997).
Extrapolation of the
RM structure function to the scattering size implies
$\delta{\rm RM}\approx 50 {\rm\ rad\ m^{-2}}$.
However, these observations are made on a much larger scale than the
scattering disk of Sgr A* and the scattering medium is believed to be
inhomogeneous.
Could the more extreme conditions necessary
to depolarize Sgr A* exist in the Galactic Center
scattering region?
The RM is expressed as
\begin{equation}
{\rm RM} = 0.8 n_e B L {\rm\ rad\ m^{-2}},
\end{equation}
where $n_e$ is the electron number density in cm$^{-3}$,
$B$ is the magnetic field parallel to the line of sight
in $\mu{\rm G}$ and $L$ is the size scale in pc.
Since $L$ must be a fraction of the scattering diameter, we find
$L \sim 0.1 \theta_{Sgr A*} D_{Sgr A*} \sim 10^{-4} {\rm\ pc}$.
Yusef-Zadeh {\it et al.\ } \markcite{yusef94} (1994) argued that the
photo-ionized skins of molecular
clouds in the GC region have a similar length scale, milliGauss fields
and $n_e \sim 10^4 {\rm cm^{-3}}$.
This matches the depolarization condition if the regions are fully
turbulent.
However, if the constraints on the outer scale of turbulence derived
by Lazio \& Cordes \markcite{lazio98} (1998) are correct, then $L \sim 10^{-7}{\rm\ pc}$.
In this case,
the RM condition and pressure balance
between the magnetic and thermal components can only be satisfied
if $B\sim 10 {\rm\ mG}$ and $n_e \sim 10^6 {\rm\ cm^{-3}}$.
These conditions are extreme even for the GC region. The largest
magnetic fields as measured for OH masers
are on the order of a few milliGauss (e.g., Yusef-Zadeh
{\it et al.\ } \markcite{yusef96} 1996). Ionized densities measured for H II regions
on the arcsecond scale ($\la 0.1$ pc) are significantly
less than $10^5 {\rm\ cm^{-3}}$
(Mehringer {\it et al.\ } \markcite{mehri93} 1993).
No depolarization is predicted for the higher temperature and lower density
model of Lazio \& Cordes \markcite{lazio98} (1998) for $B < 1 {\rm\ G}$.
The conditions necessary to depolarize at 8.4 GHz are even more extreme.
We conclude, therefore, that the conditions necessary to depolarize
Sgr A* are unlikely to occur in the scattering region.
We consider now whether depolarization may occur in
the accretion region of Sgr A*, where the electron density and
magnetic field strength are large but the length scale is smaller.
If we consider the simplest model of spherical infall with
$\dot{M}=10^{-4} M_{\sun} y^{-1}$ and equipartition between particle,
magnetic and gravitational energy (Melia \markcite{melia94} 1994),
then the change $\delta{\rm RM}$ over an interval $\delta r$ at a
radius $r$ from Sgr A* is
\begin{equation}
\delta {\rm RM} = 1.2 \times 10^{14} r^{-11/4} \delta r {\rm\ rad\ m^{-2}},
\end{equation}
where we have expressed $r$ and
$\delta r$ in units of the gravitational radius
$r_g =2 G M/c^2=7.8\times 10^{11} {\rm\ cm}$ for a $2.6 \times 10^6 M_{\sun}$
black hole.
This relation only holds for $r \ga 10^{3}$, where the temperature
falls below $10^9 {\rm\ K}$. The RM inside of this radius is
negligible unless there is a separate population of cold
electrons. We consider the effect of cold electrons in more detail in
the following Section.
If the scattering screen is at a distance of 100 pc from Sgr A*, then
the image will be an average of ray paths over a tangential length scale
$l \sim 2\times 10^{-5} r$.
Fluctuations in the RM will depolarize Sgr A*
at a given radius if $\delta{\rm RM} > 900 {\rm\ rad\ m^{-2}}$ and $l > l_0$,
where $l_0$ is the outer scale of turbulence. Assuming
$\delta r \sim r$, we find that depolarization will occur only if
$l_0 \la 10^{-5} {\rm\ pc}$. Although this scale is much smaller
than the outer scale in the local ISM (Armstrong, Rickett \& Spangler
\markcite{armst95} 1995), it is a scale that may be pertinent to the dense, energetic
environment of the accretion region.
\section{An Intrinsically Weakly Polarized Sgr A*}
A polarization fraction less than 1\% is uncommon in compact radio
sources at wavelengths shorter than 6 cm
(Aller, Aller \& Hughes \markcite{aller92} 1992). However,
optically thick
quasar cores observed with VLBI are frequently weakly polarized
(Cawthorne et al. \markcite{cawth93} 1993). Such cores may be analogous to the radio source
in Sgr A*, which, due to its low power, may not produce the strong shocks
in the jet
that are the source for higher polarization regions in quasars.
Weak polarization is
more common in radio galaxies than quasars or blazars and it is also more
common in compact-double
sources or sources with irregular morphologies (Aller, Aller \& Hughes \markcite{aller92} 1992).
A notable source with a very low polarization fraction ($<0.1\%$)
is the radio galaxy
3C 84, which has a very irregular morphology.
If the radio to
millimeter spectrum of Sgr A* does arise in a jet, the low power of this jet
or environmental effects in the Galactic Center region
may limit the magnetic field order.
For the case of a spherically symmetric emitting region, an ordered
magnetic field may depolarize the source, as well.
Alternatively, low energy electrons in the synchrotron
environment of Sgr A* may Faraday depolarize the source.
The ADAF model and the Bondi-Hoyle accretion model predict the presence
of non-relativistic electrons in the accretion region.
Observations at millimeter wavelengths may resolve many of
the questions raised in this paper. Interstellar effects are
reduced such that depolarization in the scattering region is
extremely unlikely and depolarization in the accretion region must
occur at radii less than 0.01 pc. Furthermore,
the synchrotron emission arises from a more
compact and presumably more homogeneous region. The source may also
have less synchrotron self-absorption at millimeter wavelengths,
although this is not required by all models.
We will report
in a future paper on millimeter polarimetric observations of Sgr A*.
\acknowledgements
We thank Alok Patnaik for enlightening discussions.
HF is supported by DFG grant Fa 358/1-1\&2.
|
\section{Introduction}
It has long been recognized that measurement of the profile
of the Fe K$\alpha$ fluorescence line found in many AGNs at $\sim$6.4
keV (Mushotzky 1995, Tanaka et al. 1995, Yaqoob et al. 1995, Fabian et al.
1995, Nandra
et al. 1997a and references therein) can provide an important tracer
of matter in the vicinity of the postulated
supermassive black hole.
Doppler and gravitational shifts would imprint
characteristic signatures on the line profile
which map the geometric and dynamical distributions of
matter surround the black hole. Additional information concerning the
geometry of matter in the active nucleus could in principle be derived
by studying the rapid variability of the line profile,
intensity and their relationship with the continuum variations. So far,
rapid variability in Fe K line has been detected only in two Seyfert
galaxies.
Yaqoob el al. (1996) presented the evidence for rapid variability of
the Fe K line profile in the narrow-line Seyfert galaxy NGC 7314, which is
consistent with a diskline of constant equivalent width superposed on a
constant flux narrow line (presumably from the torus).
They found while the X-ray continuum flux varied by a factor of 2 on a
time scale of hundreds of seconds, the emission in the red wing of the Fe K
line below $\sim$6 keV responded to those variations on the time scales of
less than $\sim3\times10^4$ s, and the response becomes slower and slower
towards the line peak near 6.4 keV.
Rapid variability of the Fe K line profile was also found in the bright
Seyfert 1 galaxy MCG --6-30-15 (Iwasawa et al. 1996).
On time scales of less than $\sim10^4$ s, the variability pattern is the
same as seen in NGC 7314.
But on long time scales of about several 10$^4$ s, it shows a different
behavior. When the source
is bright, the Fe K line is weak and dominated by the narrow core,
whilst during a deep minimum,
a huge red tail appears. The intensity of broad Fe K line correlates
inversely with the continuum flux, so does the equivalent width.
In this Letter we present an evidence of rapid variability of Fe K
line in another nearby (Z = 0.0023) low luminosity narrow-line
Seyfert 1 galaxy NGC 4051, which is well known for its rapid variability
in the X-ray band. Nandra \& Pounds (1994)
found an equivalent width of Fe K line $140\pm70$ eV from the Ginga
data by assuming a single narrow line. A narrow Fe K emission line was also
detected with an upper limit on the FWHM of $\sim460$ eV and equivalent
width EW $\sim170$ eV during the first ASCA observation in 1993 (Mihara el
al. 1994, hereafter M94) for this object. When the target was
re-observed by ASCA in 1994 (Guainazzi el al. 1996, hereafter G96), the
Fe K line was found to be stronger (EW = $350^{+170}_{-150}$ eV) and
broad($\sigma>0.2 keV$).
\section{The ASCA data}
NGC 4051 was observed twice by ASCA from 1993 April 25 22:30 to April 26
21:30 and from 1994 July 6 15:28 to July 9 10:40. In this
Letter, we
concentrated on the Solid-state Imaging Spectrometer (SIS) data because
of the better energy resolution it provides (Inoue 1993). The data
reductions were performed with the ASCA standard software XSELECT
according to the following criteria: satellite not passing
through the South Atlantic Anomaly, geomagnetic cutoff rigidity
greater than 6 GeVc$^{-1}$, and minimum elevation angle above Earth's
limb of $10^{\rm o}$ and $20^{\rm o}$ for nighttime and daytime
observations, respectively.
Source counts were extracted from a circular area of radius 4' for the
SIS0 and SIS1. Because of an error in satellite pointing during the first
observation, the image fell into the dead region between two chips in
SIS1 (M94). Therefore we only used SIS0 data from this observation
for spectral analysis.
To avoid the complexity caused by the "warm absorber" and "soft excess"
(M94, G96), we use the
2.5--10.0 keV data for spectral analysis. Spectral fits were carried out
using the XSPEC, and the background was taken from the blank sky data.
\section{Variability of Iron K Line}
\subsection{Comparisons between two observations}
Time average spectra of Fe K$\alpha$ line profiles for the two observations
were given by Nandra el al. (1997a) \& G96.
Due to the poor statistics, the line profile cannot be measured exactly,
therefore we model the line with a single Gaussian function.
The underlying continuum is fitted with a single
power-law absorbed by the Galactic column
density ($N_H\approx$1.3$\times$$10^{20} cm^{-2}$).
We did not take into account of the possible Compton reflection
component here because of its small impact on the ASCA spectra.
Results of the single Gaussian fits are given in Table 1.
For comparison, the results of diskline (Fabian el al. 1989, George
\& Fabian 1991) fits can be found in Nandra el al. (1997a) \& G96.
The variability of Fe K line is evident (EW changes from 166 eV to 330 eV
and FWHM from 4500 km/s to 50000 km/s), however, there is no obvious
variability in the photon index of X-ray continuum.
Because of the pointing error of the satellite during the first
observation, part of the photons fell into the gaps between the CCD
chips. The continuum flux of Obs1 in Table 1 is only a
lower limit. However, since the gaps occupy a fairly small
fraction of the selected area that includes our target ($\sim18\%$), the
continuum flux obtained cannot be biased too much. Thus we conclude
that there is no significant difference in the continuum flux
between two observations.
\subsection{Rapid FeK Line variability in the second observation}
During the two ASCA observations, NGC 4051 shows large amplitude X-ray
continuum flux changes on time scales of $\sim$100 s. In order to see if
there are short time scale variations of Fe K line within a single
observation, we choose Obs2 which last longer time ($\sim$3 days) for
analysis in detail.
Including a large flare in the beginning and a deep minimum near the end,
the whole observation has been divided into five time
intervals (from i1 to i5) with a similar exposure time ($\sim$10 ks) as
shown in Fig. 1 (different from G96).
Among the five intervals, i-2, i-3, i-4 have similar continuum fluxes.
Results of single Gaussian fittings are shown in Table 1 (We also
give the results of diskline fits to the five intervals,
respectively). The photon index
is quite similar between i2 to i4 at about $\Gamma\sim$1.80.
A steeper continuum is suggested for i-1 and a flatter one for i-5.
The single
Gaussian line dispersions and equivalent widths are also similar between
i2 to i4.
A stronger and broader line is detected in i-1, and a weaker narrower one
in i-5 (Fig. 2).
In order to show the variability of Fe K line more clearly,
we present a contour plot of Gaussian line width versus
the equivalent width for i-1, i-5 and i-2+3+4 (a summed dataset of
i-2,i-3,i-4)
obtained from the single Gaussian fits in Figure 3.
For i-1, the equivalent width of Fe K line is 733 eV, much larger than
that for the time average spectrum, and the Fe K line is much
broader ($\sigma$ = 0.96 keV). While for i-5, the EW of Fe K line is only
165 eV, and much narrower (only a narrow line has been
detected, $\sigma<$0.09 keV).
Due to the poorer statistics in i-5, there is concerning that
a weak broad Fe K component might be un-noticeable.
To exclude this possibility, we add a broad Gaussian
line (E$_{G} = 6.4$ keV, $\sigma = $0.46 keV, the same with the time
average spectrum) to the fit. The upper limit for the EW of this component
is 80 eV, far smaller than those in i-1 to i-4. These results clearly
demonstrate the large variability of Fe K line during the second ASCA
observation. G96 also analyzed the Fe K line, but failed to detect
the variability of Fe K line profiles and equivalent widths.
We would like to point out here that the time sequence
selection used in G96, different to what we did in this paper, tends to
smear out the line variations.
\section{Discussions}
\subsection{The X-ray continuum flux versus the equivalent width of Fe K
line}
We detect significant change not only in line equivalent width but also in
line profile through single Gaussian fitting to the two observations. From
1993.4.25 to 1994.7.6, Fe K line in NGC 4051 became broader and stronger,
in spite of the fact the continuum flux remains almost constantly.
This shows that the long time scale variability of Fe K line can be
independent of the continuum variation. Such kind of behavior can be
induced by the change in structure of the line emission region or the
geometry of the hard X-ray source in the context of disk line model. The
broader and stronger line emission during the 1994 epoch indicates that
the average line emission region is much closer to the putative accretion
disk.
We also detect rapid variability ($\sim10^4$s) of Fe K line within the
second observation.
During the bright flare (i-1), the Fe K line is broad and strong, while
during the deep minimum (i-5), Fe K line is narrower and weaker.
Using the ASCA data of 39 AGNs, Nandra el al. (1997b) found a clear
decrease in the strength of the Fe K$\alpha$ line with the increase
in the luminosity (X-ray Baldwin effect, Iwasawa \& Taniguchi 1993),
which was thought due to the fact that high-luminosity AGNs have high
accretion rates, causing the accretion disk to become ionized (Matt et
al. 1996).
An inverse correlation between the Fe K line EW and the X-ray continuum flux
was also found in MCG --6-30-15.
The line EWs (sum of the broad and narrow component, calculated at the
centroid energy 6.4 keV, see Table 2, Iwasawa et al. 1996) versus
continuum flux is plotted here in Fig. 4b.
The case for the NGC 4051 shows that the situation is more complicated
than this. We find an increase in the equivalent width with no apparent
change in the continuum flux when the two ASCA observations are considered.
Moreover, an increase in the line EW with the increase X-ray continuum
flux (Fig. 4a), completely opposite to the Baldwin effect and the variability
in MCG --6-30-15, is found during the second ASCA observation. It
indicates a complicate mechanism should be considered.
\subsection{rapid variability on time scale $\sim10^4$s}
It is well known that the X-ray source in many AGNs is highly variable
on short time scales $\sim$100 s. As Fe K line is produced via fluorescent
process, its strength and profile should response to the continuum
variability.
Obviously, the
measurement of the temporal response is the best way to map the matter
distribution in innermost regions around the black hole. In principle,
the distance from the X-ray continuum source to the line emission region
could be determined through observing reverberation effect (Reynolds et
al. 1998). With the Fe K line profile, we could derive the mass of the
black hole in the heart of nucleus. The characteristic time scales on
which reverberation effects occur is the light crossing time of one
gravitational radius, which is $\sim$50 s for a 10$^7$ M$_{\sun}$ black
hole. But, all current
available instruments are not able to measure such a fast time response.
Even for a bright Seyfert galaxy, ASCA can only obtain 1$\sim$2 Fe K line
photons in 100 s, too few to define the line flux and profile.
However, the Fe K line variability of time scales $\sim10^4$ s can provide
constraints on the long time scale ($\sim10^4$ s) variation of the X-ray
source or the line emission region.
So far, such variability has been detected
in only three Seyfert galaxies, NGC 7314, MCG --6-30-15 and NGC 4051. For
NGC 7314,
a simple disk-plus-torus model could explain its rapid variability
fairly well. But for MCG --6-30-15 and NGC 4051, the situation is more
complicated.
According to the disk model, the fluorescent iron lines is produced from the
hard X-ray irradiation of the disk composed of cold gas around a black hole.
When the X-ray source is bright, iron will start to be ionized, and in the
intermediate ionization states, the resonance scattering can cause a
reduction in the line flux (Matt et al. 1996). It is a possible reason
causing the anti-correlation of the EW of broad components to the X-ray
continuum flux in MCG --6-30-15. We consider that the ionized disk is also a
possible reason for the positive correlation in NGC 4051 because a factor
$\sim$2 larger EW for FeXXV than cold iron could occur (Matt et al. 1996).
Though not well constrained, the diskline energies for i1 to i5 in Table
1. show an trend to correlate with the continuum flux, which gives some
support to this opinion.
There is also another possible explanation to the the rapid variability
of Fe K line.
Fabian (1997) suggested that
the X-ray continuum might be generated by magnetic flares above the
accretion disk. At any given time there are only a few flares otherwise
the rapid variability will be average out. When the dominated flares move
around on the disk, changes in the Fe K line profile would be seen. Such
model could explain the odd variability of Fe K line in MCG --6-30-15 well.
We think it is also appropriate for the rapid variability in NGC 4051
discovered in
this paper. For the sequence i-5, the Fe K line profile is similar to the
bright flare in MCG --6-30-15. The small FWHM might due to a dominate flare
located far away from the central black hole, or a
succession of flares
on the approaching side of the disk. The broad line during i-1 might due
to a bright dominate flare very close ($\sim$6$R_g$) to the black hole, and
the large EW might due to a overabundance of iron.
\acknowledgments
This work is supported by Chinese National Natural Science Foundation,
PanDeng Project and Foundation of Ministry of Education.
The authors would also like to thank S.A. Huang for great help with
ASCA data reduction.
|
\section{Introduction}
The current interest for dwarf spheroidal galaxies (hereafter DSphs)
as satellites
of the Milky Way is raised by two fundamental questions concerning
their evolution: the real content of dark matter (DM) in these objects
(see, for example, Mateo \cite{Mateo94},\cite{Mateo97},\cite{Mateo98b};
Piatek \& Pryor \cite{Piatek95};
Burkert \cite{Burkert97} and references therein)
and their implication in the hierarchical formation of the galactic
halo (e.g., Johnston et al. \cite{Johnston96}).
The Sagittarius (Sgr) dwarf galaxy is the closest known satellite galaxy to
the center of the Milky Way, $R_{GC} \sim 16$ kpc (Ibata et al. \cite{Ibata95},
\cite{Ibata97}). Due to its proximity we can expect from its
study an additional contribution to our understanding of DSphs in general.
Since the announcement of its discovery by Ibata et al. (\cite{Ibata94})
the structure and evolution of Sgr have been extensively
discussed and simulated by various authors: Ibata et al. (\cite{Ibata95}),
Johnston et al. (\cite{Johnston95}), Vel\'azquez \& White (\cite{Velazquez95}),
Whitelock et al. (\cite{Whitelock96}),
Mateo et al. (\cite{Mateo96}), Alard (\cite{Alard96}),
Ibata et al. (\cite{Ibata97}),
Edelsohn \& Elmegreen (\cite{Edelsohn97}), Layden
\& Saragedini (\cite{Layden97}), Ibata \& Lewis
(\cite{Ibata98}), Mateo et al. (\cite{Mateo98}).
Important problems concerning this system are: 1) its possible lifetime
before dissolution, 2) the possible presence of DM in it. A
critical point in the studies mentioned above is the question whether Sgr is in
virial equilibrium or not.
\begin{table}
\caption{Observational parameters of Sagittarius dwarf galaxy (Ibata et
al. \cite{Ibata97})}
\label{Sgr_tbl}
\begin{minipage}{5cm}
\begin{tabular}{ll}
\hline
Parameter& \\
\hline
$r_{hb}$ & 0.55 kpc\\
$\sigma_o$ & 11.4 km/s \\
$\mu_{oV}$ & 25.4 mag/arcsec$^2$ \\
$L_{t}$ & $\geq 10^7$ L$_{\odot}$ \\
$(M/L)_o$ & 50 M$_{\odot}/$L$_{\odot}$ \\
$M_{t}$ & $> 10^9$ M$_{\odot}$ \\
$(l,b)$\footnote{Galactic coordinates} & (5.6$^o$, -14$^o$) \\
$(U,V,W)$\footnote{Galactic velocities} & (232,0:,194)$\pm$ 60 km/s \\
$d_{\odot}$\footnote{Heliocentric distance} & 25 kpc \\
$R_{GC}$\footnote{Galactocentric distance} & 16 kpc \\
$v_r$\footnote{Radial velocity} & 171 km s$^{-1}$\\
$(dv/db)$\footnote{Gradient of the radial velocity} & -3 km s$^{-1}$/degree \\
\end{tabular}
\end{minipage}
\end{table}
Usually, the total inferred mass of the DSph is calculated by assuming
it is in virial equilibrium\footnote{In this paper, the {\it inferred mass} is
always obtained by assuming a virial equilibrium for the satellite.}.
In this case,
the central mass-to-light ratio depends on the velocity
dispersion through the equation (Richstone \& Tremaine \cite{Richstone86})
\begin{equation}
\left(\frac{M}{L}\right)_o= \eta \frac{9 \sigma^2_o}{2 \pi G \mu_o r_{hb}}
\label{ML}
\end{equation}
\noindent
where $\eta$ is near unity for a wide variety of models, $\sigma_o$ the
central velocity dispersion, $\mu_o$ the central surface brightness and
$r_{hb}$ the half-brightness radius.
The analysis of the validity of Eq. (\ref{ML}) for evolving DSphs
has been studied in detail in the present context
by Kroupa (\cite{Kroupa97}, hereafter K97).
In Table \ref{Sgr_tbl}, we have summarized the parameters
of Sgr DSph measured by Ibata et al. (\cite{Ibata97}). For the values
of $\sigma_o$, $r_{hb}$ and $\mu_{oV}$ given in Table \ref{Sgr_tbl},
those authors obtain a high central
mass-to-luminosity ratio $(M/L)_o= 50$ M$_{\odot}/$L$_{\odot}$ by using
the Eq. (\ref{ML}). Therefore, the total inferred mass
assuming virial equilibrium, $M_{t}$, is DM dominated. The values
of total luminosity, $L_{t}$, and the total inferred mass, $M_{t}$, are
also listed in Table \ref{Sgr_tbl}.
However, the values of $(M/L)_o$ and $L_t$ suggested
by Mateo et al.
(\cite{Mateo98}) are different. These authors have discovered a tidal
extension of Sgr, which implies $L_t \leq 5.8 \times 10^7$ L$_{\odot}$ and,
as a consequence, the $(M/L)_o$ ratio could
decrease to 10 M$_{\odot}/$L$_{\odot}$.
Adopting the structural
and orbital characteristics usually assumed (given in Table \ref{Sgr_tbl}),
Ibata et al. (\cite{Ibata97})
recently suggested that such a satellite galaxy is expected to
be tidally disrupted and destroyed after several pericentric passages,
unless a significant quantity of DM is present inside it, as
inferred assuming virial equilibrium.
The observations
of the Sgr globular clusters give an age spread of its constituents $> 4$ Gyrs
and the youngest globular cluster (Terzan 7) is 9-13 Gyrs old (Montegriffo
et al. \cite{Montegriffo98}). The fast disruption obtained by the
numerical methods raise the question about the age and the
dynamical history of Sgr. However, it must be emphasized that the initial
time $t=0$
of the simulations does not necessarily coincide with the time of
formation of the oldest constituents of the satellite.
The partial formation of the
galactic halo by hierarchical processes (accretion of small galaxies) has
recently received an increase of interest, stimulated
for instance by the
investigation by Lynden-Bell \& Lynden-Bell (1995) on the reality
of streams (Lynden-Bell \cite{Lynden82})
in the close environment of the Milky Way, among them the
well known Magellanic Stream.
DSphs seem to belong to one or another of these streams and their
evolution clearly depends on their environment.
This scenario of satellite formation involves a low DM content
for the DSphs (Barnes \& Hernquist \cite{Barnes92}; Kroupa \cite{Kroupa98b}).
However,
the only simulations of no-dark matter satellite galaxies
able to survive in a tidal field are those of K97
and Klessen \& Kroupa (\cite{Klessen98}, hereafter KK98). These authors have studied a
region in the parameter space ($M_{sat}, r_{sat}$), where
$M_{sat}$ and $r_{sat}$ are the mass and a typical radius of the
satellite,
and they have obtained a
residual satellite from a more massive one.
Nevertheless, their remnant satellite
galaxies are fainter than a typical DSph, unless the true $(M/L)_{\mbox{real}} < 3$.
These circumstances make questionable the
maintenance of dynamical equilibrium and consequently the virial estimation
of mass for these systems.
In this paper we present self-consistent N-body simulations of
Sgr and a more accurate and plausible scenario of its evolution, which
suggests that Sgr is likely to be able to survive for a long time
($6-10$ Gyr)
by orbiting in the Galaxy without being dominated by DM.
In Sect. 2, we present the model of the Milky Way used in our simulations.
In Sect. 3, we describe the different models of the satellite galaxy and
corresponding scenarios of interaction with the Galaxy.
In Sect. 4 some
numerical considerations are given. In Sect. 5, we present the results of our
simulations for the different chosen
scenarios. These results are discussed in
Sect. 6 in connection with recent conjectures by Kroupa (\cite{Kroupa98},
hereafter K98) on
the parameter space possible for progenitors of
surviving DSphs without DM. Finally, a summary is given in Sect. 7.
\section{The model of the Milky Way}
\label{galaxy}
The initial model of the primary galaxy, representing the Milky~Way, is based
on the axisymmetric initial conditions adopted by Fux (\cite{Fux97})
in his N-body
modeling of the Galaxy. The mass distribution is divided in three
components: (i) an oblate stellar nucleus-spheroid with a spatial density
$\rho\propto r^{-1.8}$ in the central region, in agreement with the
near-infrared observations of the inner bulge (Becklin \& Neugebauer
\cite{Becklin68};
Matsumoto et al. \cite{Matsumoto82}),
and $\propto r^{-3.3}$ outside the bulge, as the number
density counts of stellar halo objects (e.g. Preston et al. \cite{Preston91} for
RR Lyrae; Zinn \cite{Zinn85} for globular clusters),
(ii) a double exponential stellar
disc, with radial and vertical scale lengths $h_R=2.5$~kpc and $h_z=250$~pc
and (iii) an oblate dark halo with an exponential profile ensuring a roughly
flat rotation curve out to $R=40$~kpc. Both oblate components have a
flattening $c/a=0.5$. The total mass of the luminous component is
$M_L=8.25 \times 10^{10}$ M$_{\odot}$.
The model is identical to Fux's model m04t0000, except
that the dark halo is more extended, with a scale length $b=13$~kpc and a
total (untruncated) mass $M_{\rm DH}=2.4\times 10^{11}$~M$_{\odot}$, and that
the truncation radius is moved further away at $R_{\rm c}=70$~kpc. The chosen
truncation radius ensures
a non-vanishing density everywhere along the orbit of the satellite galaxy.
The resulting rotation curve of the model is shown in Fig.~\ref{f1}.
\begin{figure}[t]
\centerline{\psfig{figure=f1.ps,width=8.8cm}}
\caption[]{Rotation curve of the initial Milky Way model (full line), with the
contributions of each component}
\label{f1}
\end{figure}
\par The initial kinematics is obtained by solving the hydrodynamical
Jeans equations, assuming an isotropic velocity dispersion for the disc
component and a anisotropic but centrally oriented velocity ellipsoid for the
other components. The velocity distribution of the nucleus-spheroid and
dark halo is generated from a 3D generalization of the Beta distribution
which limits the number of escaped particles and thus greatly improves the
equilibrium of the isolated dynamical model. More details are given in
Fux~(\cite{Fux97}).
\section{Satellite models}
The effects that a DSph suffers when it is accreted
on a primary galaxy strongly depend on its structure and
its orbit.
We have considered two kinds of satellite models. They correspond to two
scenarios of interaction between the Galaxy and the satellite.
In the first models (f-models),
we represent the satellite as a {\it standard} King's
sphere (King \cite{King66}), which matches the observational
constraints for the total
virial mass, $M_t$, the core radius, $r_o$, and the central velocity
dispersion, $\sigma_o$, of the Sgr DSph.
In this case,
the effects on the DSph are maximal because the satellite, originally
in an isolated situation, suddenly undergoes strong tidal perturbations.
Our alternative scenario (s-models) assumes that
either the DSph is formed in
the tidal tail of another major accretion event and therefore it is built
in equilibrium with the environment, or
the DSph falls slowly from a quasi-isolated situation to a tidal
field region and it has enough time to readjust itself to the environmental
forces. In both situations, we begin our simulations when the
satellite has already reached the equilibrium with the dense environment and,
therefore,
the life time of the DSph orbiting in the galaxy is expected to be longer
than in our first scenario.
In the present case, the satellite is modelled by a {\it modified} King's
sphere (G\'omez-Flechoso \& Dom\'{\i}nguez-Tenreiro \cite{Gomez97}), which takes
into account the tidal potential produced by the primary galaxy.
We have checked several orbits in order to compare the observational
features of the Sagittarius dwarf
and the numerical results in both fast and slow accreting scenarios. The
orbits we have chosen reproduce the present position, $(l,b)$, and
galactocentric velocity,
$(U,V,W)$, of
Sagittarius DSph in the observational range (Table \ref{Sgr_tbl}),
and, therefore, the simulated orbits are polar, as suggested by the
void component
of the proper motion parallel to the Galactic Plane.
The orbits are either low
eccentricity orbits in the central part of the Galaxy or high eccentricity
orbits.
The central mass-to luminosity ratio of the models has been calculated
by using the
Eq. (\ref{ML}). The central velocity dispersion $\sigma_o$ involved in this
equation has been measured along the line-of-sight. We have removed
the particles with the largest radial velocities (relative to
the radial velocity of the center of mass of the satellite) to prevent
contamination by outlying particles. We have only considered
those particles with projected distance to the center of the satellite
smaller than 0.5 kpc.
\subsection{f-models}
The distribution function of an isolated galaxy fulfills the collisionless
Boltzmann equation. We can represent an isolated DSph as a solution of this
equation. King's spheres are an example of that.
If an initially isolated DSph reaches the inner regions of
a galaxy within a short timescale,
it has no time
to modify its internal structure.
In this case the satellite maintains its isolated King's
sphere distribution function at the beginning of the simulation.
Once in the inner orbit, the satellite will
evolve quickly, because tidal forces are
strong in these regions.
This scenario is unrealistic, because,
in reality, a satellite which reaches the denser parts of the galaxy
has suffered
the influence of the galaxy potential for a long period of time.
However, we have run
simulations in such a case because they correspond to
the common initial conditions assumed in the
literature (Johnston et al. \cite{Johnston95},
Oh et al. \cite{Oh95},
Vel\'azquez \& White \cite{Velazquez95},
Johnston et al. \cite{Johnston96},
Edelsohn \& Elmegreen \cite{Edelsohn97},
K97, KK98,
Ibata \& Lewis \cite{Ibata98}) and in order to compare the results
with those of the more realistic s-model
satellites.
The parameters of the King's model we have selected for the satellite model
in a fast accreting scenario are given in Table \ref{tbl1} and they have
been chosen to reproduce the present characteristics of Sagittarius, as
was done in other simulations (Vel\'azquez \& White \cite{Velazquez95}).
The total mass of the satellite model corresponds to the virial mass
inferred for Sgr. Therefore, if the observational constraints on the
luminous mass are considered, the satellite model could represent a DM
dominated DSph.
For these models, we have assumed
$(M/L)_{\mbox{real}}=10$ M$_{\odot}/$L$_{\odot}$
for the calculations of surface brightness, which is the lower limit of the
mass-to-luminosity ratio for DM dominated satellites. A higher
$(M/L)_{\mbox{real}}$ would represent fainter satellite galaxies (for the same
DM content) and an apparently faster dissolution process.
\begin{table}
\caption{Physical parameters of the isolated King's model of Sgr (f-models):
core radius, $r_o$, central velocity dispersion, $\sigma_o$,
total mass, $M_t$, dimensionless potential, $W_o$ and tidal radius, $r_t$}
\label{tbl1}
\begin{tabular}{lllll}
\hline
$r_o$ & $\sigma_o$ & $M_t$ & $W_o$ & $r_t$ \\
(kpc) & (km/s) & ($10^7$ M$_{\odot}$)& & (kpc) \\
\hline
0.527 & 15.0 & 12.0 & 3.26 & 2.736\\
\hline
\end{tabular}
\end{table}
The initial apocenter and pericenter and the period
of the low
eccentricity orbit (f-A) and the high eccentricity orbits (f-B1 and f-B2)
are listed in Table \ref{tbl1-2}.
\begin{table}
\caption{Parameters of the orbits of the f-models}
\label{tbl1-2}
\begin{tabular}{llll}
\hline
Model & $r_{min}$ & $r_{max}$ & Period\\
& (kpc) & (kpc) & (Gyr)\\
\hline
f-A & 12 & 18 & 0.23\\
f-B1 & 8 & 38 & 0.45\\
f-B2 & 10 & 70 & 0.95\\
\hline
\end{tabular}
\end{table}
\subsection{s-models}
In the other possible scenario we assume that either the satellite
has been formed inside the tidal tail of a major merger (e.g.
numerical simulations by Barnes \& Hernquist 1992, and
the observational counterpart by Duc \& Mirabel
\cite{Duc97}) or it has been slowly accreted.
In the first case, if the dwarf galaxy is formed in equilibrium with the tidal
force of the environment, it does not contain a significant amount of
DM (Barnes \& Hernquist \cite{Barnes92}).
\begin{table}
\caption{Galaxy parameters of the s-model satellites}
\label{tbl2}
\begin{tabular}{llllll}
\\
\hline
Model&$r_o$&$\sigma_o$& $M_t$ &$W_o$&$r_t$ \\
&(kpc)& (km/s) &($10^7$ M$_{\odot}$)& & (kpc) \\
\hline
s-A & 0.06 & 11.0 & 0.93 &3.78 & 0.55 \\
s-B1 & 0.1 & 11.0 & 1.66 &3.95 & 0.99 \\
s-B2a& 0.1 & 15.0 & 5.27 &5.42 & 1.93 \\
s-B2b& 0.1 & 11.0 & 2.33 &4.87 & 1.48 \\
s-B2c& 0.3 & 15.0 & 6.04 &2.98 & 2.02 \\
\hline
\end{tabular}
\end{table}
\begin{table}
\caption{Orbital parameters of the s-model satellites}
\label{tbl2-2}
\begin{tabular}{lllll}
\\
\hline
Model&$r_{min}$&$r_{max}$&$r_{ave}$ & Period\\
& (kpc) & (kpc) & (kpc) & (Gyr)\\
\hline
s-A & 15 & 20 & 20 & 0.25\\
s-B1 & 8 & 40 & 35 & 0.45\\
s-B2 & 10 & 74 & 50 & 1.25\\
\hline
\end{tabular}
\end{table}
In the second case, according to cosmological models of hierarchical structure
formation, satellite systems are produced around massive galaxies.
These satellites could contain DM halos (e.g. Cole et al.
\cite{Cole94}).
Tidal forces could be negligible in the outer regions of the galaxy where the
satellite is formed. However, if we assume that
the satellite does not go through the denser and central regions of the
main galaxy in the first perigalacticon,
we could expect that an initially massive satellite
falls slowly on the center of the galaxy by dynamical friction, loses
part of its mass and reaches a central orbit. In this way, the
satellite has time to modify its internal structure and to reach
the equilibrium with the environment.
DSphs described in the two last scenarios have to be in equilibrium with the
tidal forces of the environment. In the paper by G\'omez-Flechoso
\& Dom\'{\i}nguez-Tenreiro (\cite{Gomez97}), the
structural parameters (total mass, $M_t$, velocity dispersion,
$\sigma_o$, core radius, $r_o$, etc.) of a satellite in the tidal
field of the primary galaxy have been estimated. Those authors have
proved that, in general, a galaxy in a tidal field can be described by
a two parameter distribution function.
They have solved altogether the Poisson equation and the
collisionless Boltzmann equation for a galaxy, taking
into account the potential of the galaxy
and the tidal potential of the environment. Only spherical terms of the
tidal field were considered in this theory.
However, the obtained {\it equilibrium} solutions are better representations
of the system than {\it isolated} models.
This result suggests that in our problem we could represent a DSph
in equilibrium with the tidal field
of the primary galaxy as a {\it modified King's sphere} (see G\'omez-Flechoso
\& Dom\'{\i}nguez-Tenreiro \cite{Gomez97}) with two free parameters.
We have chosen the central velocity dispersion,
$\sigma_o$, and the core radius, $r_o$, as input parameters, because
both can be determined from observations.
In our simulations, the values for these two parameters
are $r_o \sim 0.06-0.3$~kpc
and $\sigma_o \sim 11-15$~km/s, which reproduce the characteristics
of the Milky Way satellites. Thus, we will try to explain the
Sagittarius satellite as a typical DSph which has evolved orbiting
for a long time in the Galaxy potential.
The other parameters of the model (total mass, tidal radius and
dimensionless central potential) are automatically determined
by the tidal potential at each position of the orbit.
The satellite parameters of the models which have been performed
are listed in Table \ref{tbl2}.
The second column
is the core radius, $r_o$, and the third column is the central velocity
dispersion, $\sigma_o$, which are input parameters of the modified
King's spheres.
The total mass, $M_t$, the dimensionless central potential, $W_o$, and the
tidal radius, $r_t$, (columns 4, 5 and 6) are output parameters obtained
by solving the collisionless Boltzmann equation with the tidal potential
at the averaged distance of the orbit to the Galaxy center
(parameter $r_{ave}$ in Table \ref{tbl2-2}).
As it can be seen in Table \ref{tbl2}, the dimensionless central potential
and the mass of the satellite decrease for inner positions of the
equilibrium satellite for models with the same $r_o$ and $\sigma_o$.
Furthermore, the mass of the models is smaller
than the mass inferred from observations using kinematic arguments
(Ibata et al. \cite{Ibata97}, Mateo \cite{Mateo94}) and it is in agreement with the observed
luminous mass, assuming $(M/L)_{\mbox{real}} \sim 2-5$ M$_{\odot}/$L$_{\odot}$.
Therefore, we have assumed
$(M/L)_{\mbox{real}}=2$ M$_{\odot}/$L$_{\odot}$ for all
the s-models, that is a typical value for the
stellar population of a DSph. The low value of the mass-to-luminosity
ratio is in agreement with satellites either formed in tidal tails of major
accretion events or tidally modified by orbiting for a long time in a tidal
potential.
The apocenter, $r_{max}$, the pericenter, $r_{min}$, and the period
of the
orbits are listed in Table \ref{tbl2-2}.
The s-A orbit is an orbit of
low eccentricity in the inner region of the Galaxy and s-B1 and s-B2 orbits
have higher eccentricity.
\section{Numerical details}
\label{numerical}
The models have been evolved using the treecode algorithm kindly provided
by Barnes \& Hut (\cite{Barnes86}) with a tolerance parameter $\theta=0.7$ and
a time-step $\Delta t = 1$ Myr.
The number of particles of the luminous and dark halo components of
the primary galaxy are 15671 and 29648, respectively, and the mass
of the dark matter particles is 3 times larger that for the
luminous particles.
All the satellite models have 4000 equal-mass particles, except s-B2a
and s-B2c which have 8000 particles.
We use a softening length varying
proportionally
to the cubic root of the particle mass of the component, in order
to avoid well-known usual
numerical effects in the simulations (e.g. Merritt \cite{Merritt96}; Theis
\& Spurzem \cite{Theis99}). For
the luminous particles of the Galaxy it is $\epsilon_L =0.23$ kpc and
for the dark matter halo $\epsilon_{DH} = 0.33$ kpc. The f-models
have $\epsilon_S=0.06$ kpc, but this value is changed to
$\epsilon_S=0.05$ kpc for the s-A, s-B1 and s-B2b satellites and
$\epsilon_S=0.04$ kpc for the s-B2a and s-B2c satellites.
\section{Results}
\subsection{The main galaxy}
The main galaxy develops a bar-like structure,
described by Fux (\cite{Fux97}) for an isolated model of our
Galaxy.
The global effects of the satellite on the primary
galaxy are weak, since the mass ratio of both objects is huge. Besides,
the poor resolution of the Galaxy model prevents
a detailed description of
the local effects of the Sagittarius accretion on the Milky Way.
Therefore, we only deal with dynamical effects felt by the satellite galaxy.
\subsection{f-models}
\label{fastsat}
For our f-models, we begin
the simulations when the satellite has already reached the inner
regions of the primary galaxy.
A DSph galaxy on a low eccentricity orbit
in these inner dense regions of the Galaxy
undergoes
stronger disruptions than on more eccentric orbits,
because the tidal forces are stronger
at all positions on the trajectory.
\subsubsection{f-A satellite}
In Fig. \ref{f2}a (b), we plot the angular distribution of the
f-A satellite along (perpendicular to)
the orbit (which has eccentricity $e=0.2$)
as seen from 24 kpc away (it corresponds
to the Solar neighbourhood viewpoint), for four snapshots.
In this case, the lifetime of the dwarf galaxy,
before significant
disruption, is nearly 0.4 Gyrs
(Fig. \ref{f3}a).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f2.ps}}
\caption{f-models: mass distribution parallel
({\bf a}, {\bf c} and {\bf e}) and perpendicular
({\bf b}, {\bf d} and {\bf f}) to the
orbit as function of the angle $\phi$ to the satellite center,
as seen from 24 kpc away, which is the present distance between
the Sun and the Sagittarius DSph}
\label{f2}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f3.ps}}
\caption{Radii which enclose 70\%, 60\%, 50\%, 40\%, 30\%, 20\% and 10\%
of the initial mass of the satellite, for f-models}
\label{f3}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f4.ps}}
\caption{Aitoff projection of the f-A and f-B1 models at the
end of the simulation (2.1 Gyr)}
\label{f4}
\end{figure}
The material from the satellite galaxy is tidally stripped, forming
a long stream and then,
after 1.2 Gyrs, an almost close great circle in an Aitoff projection,
which survives for a long time (Fig. \ref{f4}a).
Moreover, as we can see in Fig. \ref{f2}b,
the mean width perpendicular to the orbit
on the sky is $7^o$ (it was only $1.9^o$ at the initial time)
and the projected surface brightness (Fig. \ref{f5}a) is
5 mag fainter at the end of the simulation (after 2.1 Gyrs).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f5.ps}}
\caption{Surface brightness $\mu_V$ of the f-models, assuming
$(M/L)_{\mbox{real},V}=10$ M$_{\odot}/$L$_{\odot}$}
\label{f5}
\end{figure}
\subsubsection{f-B1 and f-B2 satellites}
For both f-B1 and f-B2 orbits ($e=0.64$ and 0.75, respectively),
disruption mainly occurs at perigalacticon,
because
tidal forces are more efficient at small galactocentric radii.
The lifetime of a satellite depends strongly on the orbit shape.
The f-B1 satellite is destroyed after 0.5 Gyr (Fig.
\ref{f3}b) whereas
the f-B2 satellite survives for 1 Gyr (Fig. \ref{f3}c) due to the longer
period of its orbit, at this time the f-B2 satellite undergoes a close
{\it interaction} ($r \sim 2$ kpc)
with the center of the primary
galaxy and it is tidally destroyed.
The final destruction of dwarf galaxies is more efficient in
our more
eccentric orbits and they present fainter surface
brightness at the end
of the simulation than the satellite on a low eccentric orbit (Fig.
\ref{f5}). The satellite particles of f-B models are spread
over all directions and no predominant streams are formed,
as it can be seen in Fig. \ref{f4}b for the f-B1 model.
In Figs.
\ref{f2}d and \ref{f2}f,
we have represented the width (perpendicular to the orbit)
of the stream. At the end of the simulations, the f-B2 models
do not have any predominant peak in the
mass distribution perpendicularly to the initial orbit.
As a general result, due to the tidal field on the satellite
we observe:
i) a modification of the internal structure of the satellite for
both cases of orbits
(eccentric and quasi-circular), with increase of the
projected velocity dispersion, and ii) a
continuous loss of satellite mass and luminosity.
Anyone tempted to infer a $(M/L)_o$ ratio from such experiments
by assuming virial conditions would find values $100-200$ times
higher than the true ones at the disruption time (Fig. \ref{f6}),
confirming the results of K97.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f6.ps}}
\caption{Evolution of the log of
the ratio between the mass-to-luminosity
relation, $(M/L)_o$, inferred from the velocity dispersion values
under dynamical equilibrium conditions,
and the real mass-to-luminosity relation, $(M/L)_{\mbox{real}}$,
for the f-models}
\label{f6}
\end{figure}
\subsection{s-models}
\label{slowsat}
In this subsection, we analyze the interaction
effects on a DSph in equilibrium with
the galaxy potential. Either it could have fallen down slowly from an intermediate
region
in the denser parts of the primary galaxy, loosing part of its mass and
becoming a low DM satellite, or it could have been formed in a tidal
tail of a major accretion event.
\subsubsection{s-A model}
DSphs which are theoretical
equilibrium solutions to Sgr at a circular inner orbit
are small and low mass galaxies. The tidal field is almost constant
along this quasi-circular orbit and, therefore, interaction
effects on the satellites
in equilibrium are not important. For the s-A orbit, there
is a continuous loss of mass, but the satellite has still
$\sim 20 \%$ of the initial mass (Fig. \ref{f7}c)
after 2 Gyrs
and it is still detectable (Fig.
\ref{f7}a). The central surface brightness has
only changed by 2.5 mag (Fig. \ref{f7}a),
evolving from 21.3 mag arcsec$^{-2}$ to 24.8 mag arcsec$^{-2}$.
The line-of-sight velocity dispersion grows in the outer parts of the
satellite in agreement with findings by K97,
remaining almost constant in average in the inner parts ($<0.2-0.3$ kpc)
(Fig. \ref{f7}b).
We have
calculated the
mass inferred from the line-of-sight velocity dispersion, measured inside a
radius of
0.5 kpc from the satellite center, using Eq. (\ref{ML}).
Since the central velocity dispersion and the
projected density (Fig. \ref{f7}a)
evolve slowly in the central region of
the satellite at the beginning of the simulation, the inferred
mass-to-luminosity ratio
remains almost constant, contrary to the f-models. However, as
we have explained above, the real mass of the satellite decreases and
it is only $\sim 20 \%$ of the initial value after 2 Gyrs.
The structural evolution is slow (Fig.~\ref{f7}c)
and it leads to a lower central surface
brightness at the end of the simulation and a $(M/L)_o$ ratio calculated from
the velocity dispersion (inside 0.5 kpc) which is 5 times higher than
the real one (Fig.~\ref{f7}d). This effect
could be dramatically increased if the velocity dispersion is measured inside a
larger radius, because of the
velocity dispersion increase in the outer
region of the satellite, leading to a
calculated $(M/L)_o$ ratio up to 10 times larger than the real one.
Very faint tidal streams are formed, which
are spread along $\sim 75-100 \%$ of the orbit with a spatial width $\sim
6^o$ (Fig. \ref{f8}).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f7.ps}}
\caption{s-A model:
{\bf a} surface brightness $\mu_V$ assuming $(M/L)_{\mbox{real},V}=2$
M$_{\odot}/$L$_{\odot}$,
{\bf b} velocity dispersion of the satellite as a
function of the radius as same timesteps as in a),
{\bf c} evolution of the radii which enclose from 70 to 10 \% of mass
of the satellite from the top to the bottom,
and {\bf d} evolution of the mass-to-luminosity ratio
measured from the velocity dispersion. The observational
values of $\mu$, $\sigma$ are given in Table \ref{Sgr_tbl}}
\label{f7}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f8.ps}}
\caption{{\bf a} Aitoff projection of the s-A model at the end of the
simulation (2.1 Gyr), and mass distribution
{\bf b} parallel and {\bf c} perpendicular to the orbit as
function of the angle $\phi$ to the satellite center}
\label{f8}
\end{figure}
\subsubsection{s-B models}
Perigalacticon and apogalacticon of high eccentric
s-B1 and s-B2 orbits decrease with the time.
The most dramatic example is the {\bf s-B1 orbit}. The DSph passes 1 kpc
from
the galaxy center at its third passage at perigalacticon
(Fig. \ref{f9}c), because of the
orbital energy loss. During this passage
the satellite is strongly tidally stripped and it loses
most of its mass (Fig. \ref{f9}d).
The surface brightness of the satellite
decreases with time (Fig. \ref{f9}a)
and the velocity dispersion
and the mass-to-luminosity ratio increase in the outer parts
(Figs. \ref{f9}b and \ref{f9}e).
The satellite is finally completely destroyed after nearly 1 Gyr.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f9.ps}}
\caption{s-B1 model:
{\bf a} surface brightness $\mu_V$ assuming $(M/L)_{\mbox{real},V}=2$
M$_{\odot}/$L$_{\odot}$,
{\bf b} velocity dispersion of the satellite as a function of the radius,
{\bf c} distance to the satellite the Galactic Center,
{\bf d} evolution of the radii which enclose 70, 60, ..., 10\% of
the satellite mass (lines from left to right),
and {\bf e} evolution of the mass-to-luminosity ratio
measured from the velocity dispersion. The observational values
of $\mu$ and $\sigma$ are given in Table \ref{Sgr_tbl}}
\label{f9}
\end{figure}
s-B2 is
the most external and eccentric orbit which we have chosen.
As the other models, it
fulfills at certain times the position and velocity
constraints for the Sagittarius DSph orbit in
the observational range (Ibata et al. \cite{Ibata97}).
In order to investigate this case in more details,
we have selected three different satellite models for orbiting on this
trajectory (s-B2a,
s-B2b and s-B2c). The concentration (defined as $c=\log(r_t/r_o)$)
varies from 1.28 (s-B2a) to 0.83
(s-B2c). The concentration determines the fate of the satellite, as well
as the evolution of the $(M/L)_o$ ratio. The more concentrated the satellite is,
the less variation of the inferred $(M/L)_o$ ratio it suffers (Fig. \ref{f10}).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f10.ps}}
\caption{Evolution of the log of
the ratio between the mass-to-luminosity relation,
$(M/L)_o$, inferred from
the velocity dispersion values, and the real mass-to-luminosity
relation, $(M/L)_{\mbox{real}}$, for s-B2 satellites. The parameter
$c=\log(r_t/r_o)$ is the concentration of the model}
\label{f10}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f11.ps}}
\caption{Velocity dispersion, $\sigma(r)$, at four snapshots, for
s-B2 satellites}
\label{f11}
\end{figure}
The $(M/L)_o$ ratio in Fig. \ref{f10} has been calculated for
the central region of the satellite ($< 0.5$ kpc),
where the tidal effects on the mass and the velocity dispersion
are weaker. However, even in the most concentrated satellite (s-B2a), we
obtain $(M/L)_o$ ratios which are $\sim 10 (M/L)_{\mbox{real}}$, these values rise
up to several hundreds for the least concentrated model (s-B2c).
It is interesting to notice the strong variations of the
inferred $(M/L)_o$ ratio with time. This behaviour is
caused by the evolution of the surface mass density
when the satellite suffers a close encounter with the core of
the primary galaxy. It leads to a successive rearrangement of the
internal structure of the satellite.
Another important parameter in the $(M/L)_o$ calculations is the angle between
the observer-satellite line
and the main axis of the satellite. The strong
anisotropies of the satellite (tails along the orbit, anisotropic
velocity dispersion, etc) could produce various $(M/L)_o$ values (as already
suggested
by K97 and KK98).
The line-of-sight velocity dispersion evolves as shown in
Fig.~\ref{f11}.
The projected velocity dispersion decreases in the central part of the satellite
($r<0.5$ kpc) and it increases for $r>0.5$ kpc. Thus,
the $(M/L)_o$ ratio calculated from Eq.~(\ref{ML})
varies, depending on the limit radius used for the
measurement
of the {\it central} velocity dispersion
(this limit radius is related to the observational resolution).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f12.ps}}
\caption{Radii which enclose 70\%, 60\%, 50\%, 40\%, 30\%, 20\% and 10\%
of the initial mass of the satellite, for s-B2 models}
\label{f12}
\end{figure}
{\it s-B2a model:}
The {\bf s-B2a model} undergoes a long term evolution. It is the most
concentrated model that we have chosen and it survives
for at least 10 Gyrs. A mild loss of mass is produced, mainly
at perigalacticon:
the satellite looses the outer layers of mass, but a core enclosing 30\% of
the initial mass subsists (Fig. \ref{f12}a). The limit radius is smaller
than the
initial one, but the half-brightness radius increases (the final
satellite is less concentrated). The central surface brightness decreases
(Fig. \ref{f13}a), changing from $\sim 21.0$ mag arcsec$^2$ to $\sim 22.0$
mag arcsec$^2$ (assuming $(M/L)_{\mbox{real}}=2$ M$_{\odot}/$L$_{\odot}$).
The final s-B2a satellite is smaller ($r_{hb}=0.13$ kpc)
than the observed Sgr DSph ($r_{hb}=0.55$ kpc) and
the $(M/L)_o$ ratio is not large enough to reproduce the inferred $(M/L)_o
= 50$ M$_{\odot}/$L$_{\odot}$ by Ibata et al. (\cite{Ibata97}).
{\it s-B2b model:}
The mass loss of the {\bf s-B2b satellite}
(Fig. \ref{f12}b) is stronger than that of the s-B2a satellite.
The particles of the satellite outer region
are stripped, mainly
at perigalacticon. The
tidal stripped material develops a stream forward and backward from
the satellite along the orbit.
In Fig. \ref{f13}b, a low surface brightness tidal extension of
the satellite can be seen
at intermediate epochs. At the final snapshot ($t=10.5$ Gyr), the residual
core of the satellite is still detectable, it has $\sim 15 \%$ of the initial
mass (Fig. \ref{f12}b) and a central
projected surface brightness $\mu_o \sim 23.8$
mag arcsec$^{-2}$ assuming
$(M/L)_{\mbox{real}}=2$ M$_{\odot}/$L$_{\odot}$. The initial central surface brightness for the
same $(M/L)_{\mbox{real}}$ ratio was $21.3$ mag arcsec$^{-2}$. The stream formed close to the
satellite has a surface brightness which is 5.0-6.5 mag fainter than the center
of the satellite. This stream is similar to those formed by extra-tidal stars
observed close to some DSph satellites.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f13.ps}}
\caption{Surface brightness $\mu_V$ of the s-B2 satellites, assuming
$(M/L)_{\mbox{real},V}=2$ M$_{\odot}/$L$_{\odot}$}
\label{f13}
\end{figure}
At some snapshots (i.e. 3.7 Gyr, 4.83 Gyr, 6.2 Gyr, 7.3 Gyr, 8.45 Gyr
and 9.6 Gyr), when the tidal streams of the satellite are already formed,
the radial velocity, $v_r$,
and the gradient of the radial velocity, $|dv/db|$,
of the s-B2b satellite and of the
stream around the satellite reproduce the observations of Sgr
DSph region. The observational values of $v_r$ and $|dv/db|$ of Sgr
are given in Table \ref{Sgr_tbl}.
However, the values of both quantities in the models are strongly dependent on
the position and orientation of the orbit.
Therefore, small perturbations
in the orbit of the models
could lead to differences in $v_r$ of several
10 km s$^{-1}$.
In general, the sign of the variation of the radial velocity depends
on the position of the satellite along the orbit. For some particular
positions, the radial velocity gradient is almost 0 km s$^{-1}$/degree.
The s-B2b satellite at time 8.47 Gyr is an example of a model which has a
good agreement in position with the observations of Sgr DSph
(Fig. \ref{f14}c).
From the kinematical point of view, the particles in the stream
show a velocity gradient $|dv/db| \sim
4$ km s$^{-1}$/degree (Fig. \ref{f14}a), which is
slightly higher than the observed value ($|dv/db| =3$ km s$^{-1}$/degree),
whereas the radial velocity is slightly lower.
Moreover, this satellite model presents a velocity dispersion ($\sigma
\sim 6$ km/s, Fig. \ref{f14}b) lower
than the
observed value ($\sigma \sim 11.4$ km/s). However,
the projected velocity dispersion of the stream remains almost
constant and equal to the velocity dispersion
in the inner region of the system.
Observational data show that regions close to the Sagittarius globular clusters
(Terzan 8, Terzan 7, Arp 2)
have a similar projected velocity dispersion
(Ibata et al. \cite{Ibata97}).
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f14.ps}}
\caption{s-B2b satellite, at 8.47 Gyr:
{\bf a} Radial velocity, $v_r$, along the orbit
(constant galactic latitude),
{\bf b} velocity dispersion, $\sigma$, along the orbit
and
{\bf c} contour density map at the same scale as the a)
and b) plots
(the star represents the mass
center of the satellite, located at $b=-11.1^o$),
the
contours correspond to $\mu=$ 23.2, 23.6, 24.0, 24.7, 26.8 and
28.7~mag arcsec$^2$ from the center}
\label{f14}
\end{figure}
The density contour map of the s-B2b model at 8.47 Gyr (Fig. \ref{f14}c)
resembles the observational map for the region around the center of
the satellite (see for example Fig. 1 in Ibata et al. \cite{Ibata97}),
in spite of the low
resolution (few mass points) of the simulation. However, the center of our
satellite has a steeper density profile, the half-brightness
radius of the model is $r_{hb}=0.12$ kpc and that of Sagittarius DSph is
$0.55$ kpc. Even at the end of the simulation ($t=10.5$ Gyr)
the satellite remains too concentrated ($r_{hb}=0.17$ kpc). This
discrepancy could mean that
either the real Sagittarius DSph has undergone a
stronger tidal field, which has caused an effective destruction of the
satellite core, or it has suffered tidal disruption for a longer time,
or the initial density profile of the Sagittarius satellite was
more extended
than that of our model. In order to test the latter hypothesis, we
have built the s-B2c
model, which is less concentrated than s-B2b but more massive. s-B2c follows the
same orbit as s-B2b.
{\it s-B2c model:}
The {\bf s-B2c satellite} survives 5.5 Gyr before disruption. The
effects at
perigalacticon (Fig. \ref{f12}c) are stronger
than those of the other s-B2 models,
because of the lower concentration.
Low surface brightness trailing and leading streams are formed.
At 5.88 Gyr, the residual satellite looks like a long tidal extension
with a maximum projected surface brightness $\mu \sim 27.0$ mag arcsec$^{-2}$,
assuming $(M/L)_{\mbox{real}} = 2$ M$_{\odot}/$L$_{\odot}$ (Fig. \ref{f13}c).
The line-of-sight velocity dispersion evolves as shown in
Fig. \ref{f11}c.
In the outer region of the satellite
it increases up to an almost constant velocity dispersion along the orbit. At
the last snapshot of Fig. \ref{f11}c
(5.88 Gyr) the satellite is already disrupted
and the velocity dispersion of the remnant stream has increased
with respect to the value before destruction.
The central $(M/L)_o$ ratio obtained from Eq. (\ref{ML}) depends
on the position of the satellite along the orbit (Fig. \ref{f10}).
The most important variations in the $(M/L)_o$ ratio are observed at
the perigalactic passage. At this time the satellite is compressed
and stretched, its internal distribution is modified and
the tidal extensions are more pronounced. At the final steps of
the evolution,
the more stripped the satellite is, the higher measured the $(M/L)_o$ is.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f15.ps}}
\caption{s-B2c satellite, at 5.38 Gyr:
{\bf a} Radial velocity, $v_r$, along the orbit (constant galactic
latitude),
{\bf b} velocity dispersion, $\sigma$, along the orbit
and {\bf c} contour density map (the star represents the mass
center of the satellite, located at $b=-10.7^o$),
the
solid contours correspond to $\mu=24.3, 25.0, 25.5 26.7, 27.5$ mag arcsec$^{-2}$
from the center
and
the dashed contour corresponds to $\mu=30$ mag arcsec$^{-2}$}
\label{f15}
\end{figure}
We have analyzed the satellite characteristics at 5.38 Gyr from the beginning
of the simulation. It corresponds to the last approach to the galactic
disc and after the last perigalacticon, which produces total disruption. At
this time, the position ($b=-10.7^o$)
and the Galactic velocity
($(U,V,W)\sim(214,0:,170)$ km s$^{-1}$)
of the
satellite agree with the observational data. In Fig. \ref{f15},
the kinematical distribution (radial velocity
and line-of-sight velocity dispersion) along the orbit and
the contour density maps are shown. The half-brightness
radius, $r_{hb}$, and central surface brightness, $\mu_o$, of the model
roughly fit the observations (Table \ref{Sgr_tbl}).
The values we obtain in the \hbox{s-B2c} simulation at 5.38 Gyr
are $r_{hb} \sim 0.4$ kpc and $\mu_o \sim 24.3$ mag arcsec$^{-2}$, which are
slightly lower than observational values, but
they evolve to larger half-brightness radius and lower surface brightness as
the satellite approaches the Galactic disc and gets more disrupted. For
example, at $t=5.565$ Gyr we obtain the best accord between the simulations
and the observations for $r_{hb}$ and $\mu_o$ (the values of the model are
0.5 kpc and 25.4 mag arcsec$^{-2}$, respectively), but the position and
the velocity do not reproduce the observations ($b\sim 20^o$ and $(U,V,W) \sim
(103,0,86)$ km/s).
The radial velocity gradient along the orbit of s-B2c satellite depends on
the orientation of the trajectory, as in the s-B2b model. At 5.38 Gyr
we obtain $|dv/db| \sim 1.5$ km s$^{-1}$/degree for the trailing
stream, which is lower than the observed value. However, the radial velocity
$v_r$ is in agreement with the observations ($v_r=171$ km s$^{-1}$).
The satellite velocity dispersion
measured inside 0.5 kpc
has decreased
($\sigma \sim 7$ km/s), but it is still large enough to give $(M/L)_o \sim 10
(M/L)_{\mbox{real}}$.
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f16.ps}}
\caption{Surface brightness $\mu_V$ of the tidal tail close to the
Sgr satellite.
Lines represent the s-B2c model at various timesteps. Filled
squares are the data (with the errorbars)
by Mateo et al. (\cite{Mateo98})}
\label{f16}
\end{figure}
\begin{figure}
\resizebox{\hsize}{!}{\includegraphics{f17.ps}}
\caption{Tidal tail of the s-B2c satellite:
{\bf a} Radial velocity, $v_r$, along the orbit (constant galactic
latitude),
{\bf b} velocity dispersion, $\sigma$, along the orbit and
{\bf c} contour density map (the star represents the mass center
of the satellite, located at $b=5.9^o$) and the contours correspond to
$\mu = 29.6, 29.9, 30.3, 30.8, 32.0$ mag arcsec$^2$}
\label{f17}
\end{figure}
It is interesting to analyze not only the region around the center of
mass of the satellite but also the tidal tails which are formed along the
orbit. We have done it for several tidal tails close to the satellite.
In Fig. \ref{f16}, we plot
the surface brightness of the trailing tail of the s-B2c model
at various timesteps, assuming
$(M/L)_{\mbox{real}} = 2$ M$_{\odot}/$L$_{\odot}$ (a larger value of $(M/L)_{\mbox{real}}$ moves
the lines down in the figure and vice versa). The simulations are compared
with the trailing tail observed by Mateo et al. (\cite{Mateo98})
recently. As we can see, except for the region closest to the satellite
center
($<12^o$),
the simulations reproduce the observations,
giving a better result when the satellite is more disrupted. This supports
the hypothesis that Sgr DSph is close to its destruction. In our simulations,
we also obtain a leading tail similar to the trailing one. Unfortunately,
the proximity of the Galactic disc to Sgr center
makes difficult to test this result of symmetric tails
due to still poor observational data.
To give another example of the complexity of the satellite structure,
we plot in Fig. \ref{f17} the contour density levels and the
kinematic behaviour
of a tidal tail, at 5.25 Gyrs from the beginning of the simulation.
It looks like the Sgr DSph
from the kinematical point of view, having radial velocity (Fig.
\ref{f17}a) and
velocity dispersion (Fig. \ref{f17}b) similar to Sagittarius observed values
($v_r = 171$ km s$^{-1}$ and $\sigma = 11.4$ km s$^{-1}$).
The shape of the tail (Fig. \ref{f17}c) also looks
like a tidally disrupted satellite, however,
the central surface brightness is $\mu_o \sim 29.5$ mag arcsec$^{-2}$ (fainter
than a typical DSph) and the computed $(M/L)_o$ ratio is $200 (M/L)_{\mbox{real}}$
(higher than the estimation for
DSph galaxies). Nevertheless, this example illustrates,
on one hand, the difficulty of distinguishing between a tidal tails and a
DSph galaxy when the latter is close to the disruption and, on
the other hand, the
possibility of identifying objets (globular clusters, etc) which have
been tidally disrupted from a satellite galaxy by measuring their kinematic
characteristics.
\section{Discussion}
Summing up the results obtained in our simulations and comparing them
with other published works, we agree with other authors (Vel\'azquez
\& White \cite{Velazquez95}; Ibata \& Lewis \cite{Ibata98})
that Sgr DSph has a short period orbit
($T \leq 1$ Gyr). This constraint comes from measurement of the
position and the radial velocity of Sgr DSph in a realistic model of the
Galaxy.
However, we disagree about the DM content. These
previous works claim that Sgr DSph is a DM dominated satellite,
because, contrary to us, they could not obtain a low mass
satellite which survives by orbiting in the potential of the Galaxy.
Therefore, they
cannot explain the complex evolutionary pattern inferred for the
Sagittarius DSph, which has suffered a chemical
enrichment and evolution and the age interval of its globular
clusters which suggests a long life orbiting in the Galaxy.
Ibata \& Lewis (\cite{Ibata98}) have tested low massive models in order
to reproduce
the Sgr DSph characteristics, reaching an unsuccessful result on the
matter and concluding that Sgr DSph must have a large DM content.
There are two differences between their
study and ours. Firstly, they use a rigid potential to model the Galaxy, so
no energy interchange is allowed between the satellite and
the Galaxy. This restriction could eventually
prevent a readjusting of the
internal energy
of the satellite in order to reach an equilibrium with the environment.
Even if the energy transfer from the satellite to the halo is not
important in all cases, we emphasize that
our treatment is {\it a priori} adaptable to
satellites of various masses.
Secondly (and mainly),
Ibata \& Lewis (\cite{Ibata98}) do not take into account the tidal
potential of the Galaxy as they build their satellite model.
We have confirmed that
the fate of such satellites is different from that of the satellites in
equilibrium with the tidal potential as considered here.
Consequently, a more accurate model
of the initial satellite, that reflects the true dynamical situation
of the DSph, is required in order to avoid spurious effects in the simulations.
These differences in the models are responsible for the opposite conclusions
reached by Ibata \& Lewis, compared with ours concerning
DM. Furthermore, the recent observations of Sgr by Mateo et al.
(\cite{Mateo98}) prove the existence of a long trailing tail, which supports the
hypothesis that Sgr DSph has suffered strong tidal forces, it is close
to disruption and, therefore, it is not in virial equilibrium. The existence
of this tail also reduces the inferred $(M/L)_o$ ratio.
Our low DM satellites ({\it modified King's models})
establish important restrictions on the satellite formation theory. Thus,
the satellite galaxies grow either in a quasi-isolated region
(falling slowly on the center of the primary galaxy and having time for
readjusting their internal structure to the tidal forces), or inside
the main galaxy, in the tidal tails of a major accretion event, which
automatically implies equilibrium with the tidal forces at the satellite
formation epoch. Furthermore, the survival time of the {\it equilibrium}
satellites in the
potential of the Galaxy depend on the initial concentration of the DSph (the
larger concentration, the longer the life-time). We have obtain models (i.e.,
s-B2a) which survive more than 10 Gyr. The evolution of these satellites
gives rise to tidal streams and modifications on the outer velocity
dispersion of the satellite and it leads to high observed
mass-to-light ratios if dynamical equilibrium is assumed.
We have obtained a rather good qualitative
agreement with the observational
constraints for the model s-B2c in a time interval (from $5.38$ to $5.7$ Gyr),
although this agreement is not achieved simultaneously on all the
constraints.
We remind that the simulation time does not necessarily correspond with
the age of the oldest globular cluster of Sgr and, therefore, the s-B2c
model is not necessarily in disagreement with the observations.
In this case, two
possibilities could be invoked to explain the age of the oldest Sgr globular
clusters: (i) either the satellite
has been orbiting for a long time in a more external region (where tidal forces are smaller), it suffers dynamical friction or/and a deflection with
a dense structure of the MW and it reaches the present orbit where, eventually,
it is disrupted, or (ii) it has formed in a tidal
tail which contained an old stellar population (Kroupa \cite{Kroupa98c}).
The tidal tail models illustrate that a DSph without DM orbiting through
central regions of the Galactic potential could
preserve its structural parameters for a long time.
Later on, the DSph could show high inferred $(M/L)_o$ values as it becomes
disrupted. The only restriction of the model is that the satellite must
reach the central regions of the Galaxy in equilibrium with the tidal
forces. That could be possible whether the satellite has been formed in a
tidal tail of an accretion event or it has been slowly accreted
from the external parts of the Galaxy. The latter possibility could increased
the estimated life-time of the satellite before destruction.
Recently, K98 has discussed the parameter space of the
DSphs which could survive in a tidal field. In spite of the
different initial distribution functions, our s-models
seem to be
located in the available region of the ($M_{sat}$, $r_{sat}$) space for
surviving satellites. However, the main difference between
the approaches to the problem is that we build the model of the
satellite initially in equilibrium with the tidal forces by solving the
Boltzmann equation (G\'omez-Flechoso \& Dom\'{\i}nguez-Tenreiro
\cite{Gomez97}) and that assures {\it a priori} the survival of the satellite
for a long time.
\section{Summary}
We have performed simulations of Sagittarius-like satellites to follow
the evolution of these systems.
The satellite and the primary galaxy are modelled as N-body
systems, that prevents possible purely numerical effects which could appear
in a non-selfgraviting scheme (i.e. if one of the interacting
system is represented by a rigid potential). Two satellite models
have been tested:
\begin{enumerate}
\item First of all, we have used {\it isolated King's models} (three
free parameter models) for the satellite, but they are tidally stripped and
destroyed in a short interval of time, undergoing a fast evolution.
\item In order to correctly model
the observed situation we had to introduce the
{\it modified King's models}
(two free parameters-models),
which allow us to build up the distribution
function of the satellite, taking into account the tidal force of the
environment. These s-models could correspond to no DM dominated satellites.
In this framework,
the initial concentration of the satellite
and the trajectory determine its life-time. Central circular orbits lead to
more effective tidal destruction events than highly eccentric orbits.
For a given
kind of orbit, the most concentrated satellites live longer.
The model
which better reproduces the observations is the s-B2c model after
orbiting 5.38-5.6 Gyr in the potential of the Milky Way.
For the same orbits, the s-models survive a longer time orbiting
in the Galactic potential than the f-models, despite the
s-models having smaller masses (smaller binding energies).
\end{enumerate}
According to the results presented here,
our
main conclusions about the evolution of Sgr are:
\begin{enumerate}
\item Sagittarius
may not be
a dominated DM satellite.
\item It follows an eccentric orbit (perigalacticon and apogalacticon
are approximately 15 and 70 kpc).
\item It was more concentrated in the past than
at the present epoch.
\item It could have
been orbiting in the Galactic halo for a long
time (minimum 5 Gyr).
\end{enumerate}
As general conclusions about the DSph satellites, we can remark that:
\begin{enumerate}
\item High values of $(M/L)_o$ found in the literature (Ibata et al. \cite{Ibata95};
Irwin \& Hatzidimitriou \cite{Irwin95}; Mateo \cite{Mateo98b}
for a review of the Galactic
satellites) could be due to an erroneous use of the
virial theorem and not to the presence of DM (as
already suggested by K97;
KK98).
\item The progenitor of these satellite could be either lumps formed in a
major accretion event or other dwarf galaxies (maybe dwarf irregulars or
dwarf spirals) which fall in the
potential well of the Galaxy, lose their gas content and evolve to the
equilibrium configuration in the tidal field of the Milky Way, giving rise
to the population of DSphs. The finding of intermediate states of this
evolution could be a support to the proposed scenario.
\item According to some of the present results, some DSphs are able to
survive for long times orbiting in the Galactic halo.
That should be taken into account for the calculations of the dissolution
rate of dwarf galaxies
in the halo of the primary galaxy.
\end{enumerate}
\begin{acknowledgements}
We would like to thank D. Pfenniger and D. Friedli for critically reading
the paper, as well as the referee for remarks which have led to
clarify various points of the paper.
M.A. G\'omez-Flechoso was supported by the Fundaci\'on Ram\'on Areces
through a fellowship. This work has been partially supported by the Swiss
National Science Foundation.
\end{acknowledgements}
|
\section{Introduction}
Brane constructions in string theory provide powerful tools for analyzing
field theories in diverse dimensions and with varying
amounts of supersymmetry~\cite{branes1,branes2}. For a review and
references see \cite{review}. More recently the Maldacena conjecture
\cite{m1,gubs,w1} added a new relation between the large $N$
limit of conformal field theories on branes and the near horizon geometry of
the corresponding black brane solutions. The original conjecture was stated
for ${\cal N}=4$ SYM realized on $N$ 3-branes, but subsequently more general
examples were discovered. One class of such examples are orbifolds of the
${\cal N} = 4$ configuration \cite{ks,oz} and another includes theories on
3-branes in nontrivial F-theory backgrounds \cite{f1,f2,kap}. All of these
constructions give rise to conformal theories with varying amounts of
supersymmetry. A third class of theories arises on 3-branes at a conifold
singularity \cite{kw,ur,nek,unge,josh}. These ${\cal N}=1$ theories are not
conformal at all scales, but flow to a line of conformal fixed points in the
infrared. For all these theories the correspondence between the large $N$
field theory and supergravity was studied in some detail.
For branes on a conifold it turned out to be useful to have a type IIA description which is
related to the IIB configuration via T-duality~\cite{ur}.
In this paper we study a superconformal ${\cal N} = 1$ $Sp(N)\times Sp(N)$
gauge theory with matter in the fundamental, bifundamental, and antisymmetric
representations. We also discuss a specific
deformation which preserves ${\cal N}=1$ SUSY but breaks conformal invariance.
The resulting theory has a running gauge coupling and flows to a line of
superconformal fixed points in the infrared.
For both of these theories we give a IIA brane construction as well as a IIB
orientifold construction. The latter description allows us to obtain the supergravity
solution that is dual to the large $N$ limit of the conformal field theory.
The type IIA description, on the other hand, provides a simple way to determine
the gauge group, the matter content, and the superpotential of the theories in
question.
In most ${\cal N}=1$ theories discussed in the AdS/CFT literature (see e.g.
\cite{ks,kw})
the R-current which is the superpartner of the stress-energy tensor can be
fixed uniquely
by field theory considerations. For the theories we discuss here this is not
the case. There
is a one parameter family of candidate R-currents, both in the theory with
vanishing beta function, and in its deformation which flows to a line of fixed
points.
Since the R-charges of the fields are not uniquely determined, there is no
field theory prediction for the
dimensions of the chiral primary operators. On the other hand, once we have a
supergravity dual of the large $N$ field theory, we know which gauge boson on
AdS is the superpartner of the graviton. If we are able to match
field theory operators
with supergravity states, we can determine the R-charges
of all fields and therefore the dimensions of all chiral primary operators.
Although there is no firm field-theoretical prediction for the dimensions of fields
in the infrared, for the theory with vanishing beta function the most natural
assumption is that all fields have canonical dimensions, i.e., that the theory
is finite. This will
be born out by the supergravity analysis. In the other case, the theory with
a running coupling constant, the correct charge assignment in the
infrared is harder to guess. Unfortunately the supergravity analysis in this case is
on a considerably less solid footing and depends on circumstantial evidence.
Nonetheless our analysis suggests a definite R-charge assignment. It would be interesting
to find a field theory explanation for it.
The type IIA construction involves D4-branes compactified on a circle
as well as NS5-branes, D6-branes, and O6-planes. The gauge theory lives
on the D4-branes. Our construction is very similar to the
brane configurations that give rise to elliptic ${\cal N}=2$ models
\cite{branes2,angfinite,jaemo,kap}. One advantage of the IIA description is that the
moduli space of the gauge theory is realized geometrically. The flat directions
correspond to motions of the 4-branes. Similarly, relevant perturbations
of the field theory, such as masses for the matter fields, are also realized
geometrically as motions of the 6-branes. This allows us
to identify a 6-brane configuration that gives rise to a superconformal
${\cal N}=1$ theory on the 4-branes with an exactly marginal parameter.
We can also identify relevant perturbations of the
superconformal 4-brane theory that lead to theories with running coupling
constants. There is one particular perturbation that gives rise to a theory
that flows to a line of conformal fixed points in the infrared.
The moduli space of the perturbed theory has a Coulomb branch. A generic
${\cal N} = 1$ theory with a Coulomb branch has a low
energy effective gauge coupling that varies over the moduli space. The theory
we are considering in this paper has the special feature that the low-energy effective
gauge coupling does not depend on the moduli. This will be relevant
when we discuss the supergravity description of these theories.
In order to construct the supergravity duals we
T-dualize the IIA configuration along the compact direction. This operation
turns the D6-branes and the O6-planes into D7-branes and O7-planes. The D4-branes
turn into D3-branes probing this background. Similar probe theories were
studied
in \cite{bds,jhs,aks,ah1}, and their relation to supergravity is described in
\cite{f1,f2,kap}. Our IIA configuration turns out to be T-dual to 3-branes
probing a local
piece of an F-theory compactification \cite{sen1,sen2} which is related to
the Gimon-Polchinski model \cite{gp}. The simplest such configuration,
consisting of two intersecting O7-planes with four coincident 7-branes
on top of each, corresponds to the IIA construction of the superconformal
4-brane theory. In the type IIB construction the Ramond-Ramond (RR) charges of the
7-branes are cancelled locally by the charges of the orientifold planes, so the string
coupling is constant. Since the type IIB description is a perturbative orientifold,
we can find the supergravity dual of the large $N$ limit of the field theory
along the lines of \cite{f1,f2}. Matching the spectrum of primary operators
with the KK modes allows us to determine the $U(1)_R$ charges of
all fields in the conformal theory unambiguously. The matching of non-chiral
primaries exhibits a new interesting feature: we find a short supergravity
multiplet whose field theory counterpart becomes short only when $N\to\infty$.
We interpret this as the evidence that at higher orders in $1/N$ supersymmetry
mixes one-particle and two-particle supergravity states.
It should also be possible to find a supergravity description of the
infrared limit of the deformed theory. Although this theory is not conformal, it
has a constant low-energy
effective coupling along the Coulomb branch, so the supergravity dual will have a
constant dilaton. To find this dual we need to study the
deformations of the backgrounds in IIA and IIB and find an explicit map between them.
As mentioned before, this is straightforward on the IIA side, since the deformations
correspond to motions of the 6-branes. On the IIB side the situation is more
involved. We can analyze the deformations on the IIB side by studying the
theory on the 7-branes.
The eight-dimensional theory on the 7-branes has six-dimensional matter
localized at the intersection of orthogonal 7-branes. We analyze the moduli
space of this impurity theory following \cite{kap2}, and find an
explicit map between the type IIB and type IIA deformations.
Among supersymmetric type IIB deformations there is one that maps to a new IIA
brane configuration which involves curving D6-branes in the background of an
NS5-brane. The map between deformations also allows us to identify the IIB
configuration that gives rise to the non-conformal probe theory
with moduli-independent effective coupling. We do not have a complete supergravity
description of this theory, but a partial description is possible. It supplies enough
information to determine the dimension of all chiral operators in the infrared if
we use field theory considerations as well.
In section \ref{IIA} we discuss the type IIA construction of the probe theory
and list some field theory results that we need in subsequent sections.
Section \ref{IIB} contains the T-duality, the analysis of the 7-brane
impurity theory, and the map between IIA and IIB deformations. We also
briefly discuss the exotic IIA deformation that appears as the counterpart of
an ordinary deformation in IIB. In section \ref{sugra} we analyze the large $N$
limit of our field theories and their supergravity duals. We discuss the matching
of operators with Kaluza-Klein modes in the conformal case and present a
partial analysis in the non-conformal case.
\section{The IIA construction of the field theory}
\label{IIA}
\subsection{The IIA brane configuration}
A configuration consisting of D4-branes extending in 01236, D6-branes and
O6-planes extending
in 0123789, and NS5-branes extending in 012389 preserves four supercharges.
We obtain an
${\cal N}=1$ supersymmetric field theory in four dimensions after compactifying
$X_6$ on a circle with circumference $2\pi R_6$. Specifically we consider
configurations
with $N$ D4-branes wrapping the compact $X_6$ direction. We put two
O6$^-$-planes at
$X_6=0,\pi R_6$ and an NS5-brane and its image at $X_6 = R_6\pi/2,3R_6\pi/2$. In
order to cancel
the total RR charge, we place four physical D6-branes on the circle.
An example of such a configuration is shown in Fig.~\ref{fig1}.
These brane configurations are very similar to the configurations that give
rise to finite
${\cal N} =2$ theories in four dimensions \cite{angfinite,jaemo,kap}.
In fact, the configuration we study here can be obtained from one of the ${\cal N} =2$
configurations in \cite{angfinite} by rotating the NS5-branes from the 45 directions into
the 89 directions. This breaks half of the supersymmetries, giving an ${\cal N}=1$
theory in $d=4$.
\begin{figure}
\centerline{\epsfxsize = 11truecm \epsfbox{fig1.eps} }
\tighten{
\caption{ Brane configuration: The vertical dashed lines are the
O6-planes, the solid lines are the D6-branes, the horizontal line are the
D4-branes, and the point represents the NS5-brane. Only half of the $X_6$
circle is shown.}
\label{fig1}
}
\end{figure}
Using standard techniques \cite{review}, we can determine the matter content
and the superpotential of the field theory on the 4-branes.
Unlike the ${\cal N} =2$ case, the $X_6$ position of the D6-branes will
play an important role in our analysis. We need to distinguish two cases
that are of interest for the analysis in this paper.
Either all 6-branes intersect the NS5-brane,
or the four 6-branes are split into two groups of two to the left and right
of the NS5-brane (as shown in Fig.~\ref{fig1}). These two choices give rise
to physically inequivalent theories. The former configuration yields a line
of fixed points (parametrized by the dilaton expectation value) that passes
through zero coupling,
while the latter corresponds to a non-conformal gauge theory which flows to
a line of strongly coupled fixed points.
\subsection{The conformal case: a field theory analysis}\label{conformal}
\label{iib}
The theory on the 4-branes turns out to be an $Sp(2N)_1\times Sp(2N)_2$ gauge
theory with matter fields $A_i, i=1,2$ in the antisymmetric representation
of each of the gauge groups, two bifundamentals ${\cal Q}, \tilde{\cal Q}$,
and fundamentals from the 4-6 strings. The brane configuration, and consequently
the field theory, admit a symmetry which exchanges the two $Sp$ factors. To
determine the number and flavor
representations of the fundamentals we need to understand the classical
gauge theory on the 6-branes. Note that the worldvolume of the NS5-brane
lies within the worldvolume of the D6-branes. It was argued in \cite{karch}
that the 6-branes can break on the NS5-brane (see also \cite{nscft}). The gauge
group on the four 6-branes turns out to be $U(4)_u\times U(4)_d$, where the
two $U(4)$ factors act on the upper and lower halfs of the 6-branes respectively.
One-loop effects break the $U(4)_u\times U(4)_d$ symmetry to
$SU(4)_u\times SU(4)_d$~\cite{berkoozetal}.
The matter content of the 6-brane theory includes a bifundamental
hypermultiplet from strings connecting upper and lower halfs of the 6-branes.
We will have more to say about the 6-brane theory when we discuss the
deformations of this background. For our present purposes we only need to
know that the gauge group of
the 6-brane theory is the flavor group of the probe theory.
The matter content and the superpotential
for a 4-brane probe in this background were worked out in \cite{karch}.
The fundamentals transform as $q=(\Yfund,1,{\bf 4},1)$,
$\tilde{q}=(\Yfund,1,1,{\bf 4})$,
$p=(1,\Yfund,{\bar {\bf 4}},1)$, and $\tilde{p}=(1,\Yfund,1,{\bar {\bf 4}})$
under
$Sp(2N)_1\times Sp(2N)_2\times SU(4)_u \times SU(4)_d$. The superpotential
reads
\begin{equation}
W = h_1\tilde{\cal Q}A_1 J_1{\cal Q} - h_1{\cal Q}A_2 J_2\tilde{\cal Q} +
h_2q{\cal Q} p + h_2\tilde{p}\tilde{\cal Q}\tilde{q}.
\end{equation}
Here $J_1$ ($J_2$) is the invariant antisymmetric tensor of $Sp(2N)_1$ ($Sp(2N)_2$).
Following \cite{line} it is easy to check that this theory has a line of
fixed points passing through weak coupling. The one-loop beta function vanishes
and the symmetry between the gauge factors implies that both antisymmetric
tensors have the same anomalous dimension $\gamma_A$, both bifundamentals
have $\gamma_{\cal Q}$ and all fundamentals have $\gamma_q$. Therefore
the beta functions of the gauge coupling and the
Yukawa couplings in the superpotential are
\begin{eqnarray}\label{beta1}
\beta_g &\sim& 2(N-1)\gamma_A+4N\gamma_{\cal Q}+8\gamma_q \nonumber \\
\beta_{h_1}&\sim& 2\gamma_{\cal Q}+\gamma_A\\
\beta_{h_2}&\sim& \gamma_{\cal Q}+2\gamma_q. \nonumber
\end{eqnarray}
Setting all beta functions to zero gives two independent constraints on the
three coupling constants. The remaining coupling constant parametrizes a
line of superconformal fixed points. Since setting all anomalous dimensions
to zero satisfies the constraints, this line passes through the free point
$g=h_1=h_2=0$. Note that requiring the beta functions to vanish does not
fix anomalous dimensions unambiguously. The most natural assumption is
that the dimensions of the fields are unchanged as one moves along the fixed
line. This would mean that the theory is finite.
The supergravity computation in the last section supports this
conjecture by showing that this is true in the large $N$ limit.
The moduli space of this theory includes subspaces where it flows to
theories with more supersymmetry. For example, giving an expectation value
to either ${\cal Q}$ or $\tilde{\cal Q}$ proportional to a unit matrix gives a
mass to half of the fundamentals and breaks the gauge group to the diagonal
$Sp(2N)_D$. It is a simple matter to show that the resulting theory flows to an
${\cal N}=2$ superconformal theory with gauge group $Sp(2N)$, one antisymmetric
hypermultiplet, and four hypermultiplets in the fundamental. Giving such
expectation values to both ${\cal Q}$ and $\tilde{\cal Q}$ makes all flavors massive
and breaks the gauge group to $SU(N)$. Part of the bifundamentals are eaten by gauge
bosons, and the rest give rise to three chiral superfields in the adjoint of $SU(N)$.
This theory flows to ${\cal N} = 4$ SYM in the infrared.
These field theory results are reproduced in the brane construction
if we identify the positions of the D4-branes with the field theory moduli
in the following way:
\begin{eqnarray}
&X_7 \sim {\cal Q Q}^\dagger - \tilde{\cal Q}^\dagger \tilde{\cal Q} & \\
&X_4+iX_5 \sim {\cal Q}\tilde{\cal Q}.&
\end{eqnarray}
Giving an expectation value to either of the bifundamentals while keeping the other
expectation value zero
corresponds to moving the 4-branes in the positive or negative $X_7$ direction.
Turning on both bifundamentals corresponds to moving the 4-branes in the $X_4$
and $X_5$ directions as well as $X_7$. The effect of these motions on the 4-brane theory
agrees with the field theory expectations. If we move the 4-branes off the NS5-branes in
the $X_7$ direction, we can ignore the NS5-brane. The remaining branes preserve eight
supercharges, and standard techniques \cite{review} confirm the matter content and gauge
group stated above for the ${\cal N}=2$ case. Moving the 4-branes in $X_4$
and $X_5$ amounts to separating them from all other branes. The theory on
the 4-branes is then ${\cal N} = 4$ SYM as expected.
\subsection{The non-conformal case}\label{iic}
We can deform the background for the 4-brane theory by moving the 6-branes
in the $X_6$ and $X_{4,5}$ directions. These brane motions are parametrized
by expectation values of the two complex scalars, ${\cal M}, \tilde{\cal M}$
in the bifundamental hypermultiplet of the $(1,0)$ theory on the intersection of
the 6-branes and the NS5-branes \cite{karch}.
More precisely, we relate the positions of the 6-branes to
${\cal M}, \tilde{\cal M}$ as follows:
\begin{eqnarray}
&X_6 \sim {\cal M M}^\dagger - \tilde{\cal M}^\dagger \tilde{\cal M} & \\
&X_4+iX_5 \sim {\cal M}\tilde{\cal M}.&
\end{eqnarray}
To obtain the configuration shown in Fig.~\ref{fig1} we have to set
${\cal M}={\rm diag}(m_1,m_2,0,0)$ and
$\tilde{\cal M}={\rm diag}(0,0,\tilde{m}_3,\tilde{m}_4)$.
These bifundamental expectation values act as mass terms in the 4-brane theory.
The corresponding terms in the field theory superpotential are
\begin{equation}
W = \tilde{\cal M}q\tilde{q} + {\cal M}p\tilde{p}.
\end{equation}
We will be particularly interested in the case
$m_1=m_2=\tilde{m}_3=\tilde{m_4}$.
In this case the bifundamental expectation values break the
$SU(4)_u\times SU(4)_d$ 6-brane gauge group
to $SU(2)_1\times SU(2)_2\times U(1)$.
After integrating out the massive components of the fundamentals, the
superpotential of the 4-brane theory reads
\begin{equation}\label{defW}
W = h_1\tilde{\cal Q}A_1 J_1{\cal Q} - h_1{\cal Q}A_2 J_2\tilde{\cal Q} +
h_3 q{\cal Q}\tilde{\cal Q}\tilde{q} + h_3 \tilde{p}\tilde{\cal Q}{\cal Q}p.
\end{equation}
The fundamentals now transform as $q=(\Yfund,{\bf 1},{\bf 2},{\bf 1})$,
$\tilde{q}=(\Yfund,{\bf 1},{\bf 2},{\bf 1})$, $p=({\bf 1},\Yfund,{\bf 1},{\bf 2})$, and
$\tilde{p}=({\bf 1},\Yfund,{\bf 1},{\bf 2})$ under
$Sp(2N)_1\times Sp(2N)_2\times SU(2)_1 \times SU(2)_2$. Actually, the
superpotential, \eq{defW}, has an accidental $SO(4)_1\times SO(4)_2$ global symmetry
under which $q$ and $\tilde{q}$ transform as a $({\bf 4}, {\bf 1})$ while
$p$ and $\tilde{p}$ transform as $( {\bf 1},{\bf 4})$.
An analysis along the lines of \cite{line} shows that this theory also has a
line of superconformal fixed points. The beta functions are given by
\begin{eqnarray}\label{beta2}
\beta_g &\sim& 4+2(N-1)\gamma_A+4N\gamma_{\cal Q}+4\gamma_q \nonumber \\
\beta_{h_1} &\sim& 2\gamma_{\cal Q}+\gamma_A\\
\beta_{h_3} &\sim& 1+\frac{1}{2}\gamma_{\cal Q}+\gamma_q. \nonumber
\end{eqnarray}
Demanding that the beta functions vanish, we again find that two out of the
three constraints are independent, leaving us with a line of fixed points.
In this case, however, the line does not pass through weak coupling, since at
least one of the anomalous dimensions must be nonzero. Again the vanishing of
the beta functions alone does not determine the values of anomalous dimensions.
In the last section we will argue that supergravity considerations allow us
to fix this ambiguity for large $N$ and find $\gamma_A=\gamma_Q=0,\gamma_q=-1$.
As in the conformal case we can analyze the RG flows both in field
theory and using the brane picture. From the brane construction it is clear
that we flow to the same ${\cal N}=2$ theory as in the conformal case if we move
the 4-branes off the NS5-brane in the positive or negative $X_7$ direction. Moving the
4-branes in $X_{4,5}$ again yields ${\cal N}=4$ SYM. The analysis in the field
theory is a little more involved in this case because the one-loop
beta function does not vanish. This implies that there will be threshold
effects in the matching of the running gauge coupling. On general grounds
one would expect the low-energy effective coupling to depend on the size of the
bifundamental expectation values in the field theory. However, if we give arbitrary
(non zero) expectation values to ${\cal Q}$ and $\tilde{\cal Q}$, fields get integrated
out at a variety of scales. Assuming that the expectation value of ${\cal Q}$ is larger
than that of $\tilde{\cal Q}$, the $Sp(2N)\times Sp(2N)$ gauge group is broken to the
diagonal group at a scale set by ${\cal Q}$. The diagonal $Sp(2N)_D$ is
broken to $SU(N)$ at a scale set by $\tilde{\cal Q}$, and finally the
fundamentals are integrated out at scale $h_3 {\cal Q}\tilde{\cal Q}$. Matching
the gauge couplings at each of these scales we find that the low-energy
effective coupling does not depend on the bifundamental expectation values. This is a
special feature of this theory that will be important later on.
\section{The type IIB description}
\label{IIB}
\subsection{T-duality}\label{iia}
In this section we describe the IIB configuration which is obtained by T-dualizing
the IIA brane configuration of section II along $X_6$. Since $\partial /\partial X_6$
is not a Killing vector, performing this T-duality is not completely trivial.
Similar T-dualities on IIA configurations that
preserve ${\cal N}=2$ supersymmetry on the 4-branes have appeared in the
literature \cite{kap,jaemo}. In the ${\cal N}=2$ case the T-duality maps
the two O6$^-$-planes and the four D6-branes to an orientifold 7-plane and four
D7-branes. The D4-branes become D3-branes probing this background. The NS5-brane
and its mirror image turn into a ${\bf Z}_2$ orbifold acting on the 7-brane coordinates
transverse to the D3-brane. The T-dual of the IIA configuration without NS5-branes was
analyzed in \cite{bds,ah1}.
Our configuration differs from the ${\cal N}=2$ case by the orientation of the
NS5-branes. Since this modifies the T-duality considerably we discuss it in
some detail here.
Our first goal is to T-dualize the NS5-branes and the pair of O6$^-$-planes.
The other branes can be added later.
We begin by separating the NS5-brane and its image in the $X_{4,5}$ directions.
The T-dual of the two NS5-branes is a two-center Taub-NUT space.
Recall that the two-center Taub-NUT space can be thought of as a circle fibered
over ${\bf R}^3$ so that its radius vanishes at two points on ${\bf R}^3$ (the
centers). In the present case ${\bf R}^3$ is parametrized by $X_4,X_5,X_7$, while
the coordinate along the circle is T-dual to $X_6$. The positions of the centers
correspond to the positions of the NS5-branes in $X_4,X_5,X_7$.
In the IIA configuration the orientifold projection ensures that position of the physical
NS5-brane and its image are related by a reflection of the $X_{4,5}$ coordinates.
The T-dual orientifold projection should therefore impose a similar
constraint on the location of the centers of the Taub-NUT.
The Taub-NUT metric has the following form
\begin{eqnarray}
ds^2 &=& \left(\frac{4}{b^2}+\frac{1}{R_+}+\frac{1}{R_-}\right)^{-1}
\left[ d\sigma+
\left(\frac{Z_+}{R_+}+\frac{Z_-}{R_-}\right)d\arctan\left(\frac{Y}{X}\right)
\right]^2 \nonumber \\
&&+\left(\frac{4}{b^2}+\frac{1}{R_+}+\frac{1}{R_-}\right)
\left[ dX^2+dY^2+dZ^2\right],
\end{eqnarray}
where
\begin{equation}
Z_\pm = Z \pm Z_0 \qquad R_\pm^2 = X^2+Y^2+Z^2_\pm.
\end{equation}
The $\bf{R}^3$ base is parametrized by $X,Y,Z$, the two centers are located at
$(0,0,\pm Z_0)$, and $\sigma$ is the $4\pi$-periodic coordinate on the circle fiber.
The parameter $b$ is the asymptotic radius of the fiber. The reflection of
$X_{4,5}$ in the IIA picture map into reflections of $Z$ and one other
coordinate of $\bf{R}^3$, say $Y$.
We will be interested in the limit when the asymptotic radius of the circle
fiber, $b$,
becomes infinitely large, while the T-dual circle parametrized by $X_6$ shrinks
to zero. In this limit the two-center Taub-NUT space becomes an $A_1$ ALE
space, also known as Eguchi-Hanson space. It is useful to change coordinates
\cite{prasad} to transform the metric above into the Eguchi-Hanson form:
\begin{eqnarray}
X &=& \frac{1}{8} \sqrt{r^4-a^4} \sin(\theta)\cos(\psi) \\
Y &=& \frac{1}{8} \sqrt{r^4-a^4} \sin(\theta)\sin(\psi) \\
Z &=& \frac{1}{8} r^2 \cos(\theta)\\
\sigma &=& 2\phi,
\end{eqnarray}
where $a^2=8Z_0$ and $\psi$ has period $2\pi$.
The orientifold-induced projection $(Y,Z)\sim (-Y,-Z)$, implies the
identification
$(\theta,\psi)\sim (\pi-\theta,-\psi)$ for the angular coordinates. The fixed
locus of this identification is a two-dimensional submanifold of the Eguchi-Hanson
space which has the topology of a cylinder. Next we want to bring
the NS5-brane and its image back to the origin of
the $X_{4,5}$ plane in the IIA description, which corresponds to setting $a=0$.
For $a=0$ the Eguchi-Hanson metric becomes an orbifold metric on ${\bf C}^2/{\bf Z}_2$.
To make this explicit we can introduce two complex coordinates
\begin{equation}\label{z12}
z_{1,2} =
r\exp(i\phi/2)\left(\cos(\theta/2)\exp(i\psi/2)\pm
i\sin(\theta/2)\exp(-i\psi/2)\right).
\end{equation}
In these coordinates the $a=0$ Eguchi-Hanson metric becomes flat.
The identification $\psi\to \psi+2\pi$ requires that we identify $(z_1,z_2)
\to (-z_1,-z_2)$ as expected for ${\bf C}^2/{\bf Z}_2$. The additional
orientifold identification acts on the new coordinates as
$(z_1,z_2)\to(z_1,-z_2)$, and acting with both orientifold and orbifold
identifications flips the sign of $z_1$. The orientifold projections have two
fixed planes, $z_{1,2}=0$, which we identify with
two O7$^-$-planes. To summarize, the NS5-brane together with two O6$^-$-planes
become, under T-duality, a pair of intersecting O7$^-$-planes with six common directions.
Now let us put in D-branes. The four physical D6-branes in IIA are located at
$X_4=X_5=0$. Under T-duality they become D7-branes wrapping the circle fiber of
the Taub-NUT and located at $Y=Z=0$. In other words, they are wrapped on the
invariant cylinder of the orientifold projection. Taking the limit
$b\to \infty, a\to 0$ we
find that the invariant cylinder develops a neck and becomes a pair of planes
$z_1=0$ and $z_2=0$ in ${\bf C}^2/{\bf Z}_2$. Thus the four physical D7-branes must be
located on these planes. Recall that these planes are the O7$^-$-planes and
therefore have 7-brane charge $-4$. It follows that the 7-brane charge is
cancelled between the D7-branes and the orientifold planes,
and the IIB dilaton is constant. Finally, T-duality turns the D4-branes into
D3-branes extending in 0123. To summarize, the T-dual of the IIA
configuration in the limit
when the radius of $X_6$ goes to zero consists of an O7$^-$-plane
with four coincident D7-branes in 01236789, another O7$^-$-plane with
four coincident D7-branes in 01234589 and 3-branes in 0123. We will refer to
the 7-branes extending in 01234589 as 7$'$-branes. The orientifold group for
this configuration is
\begin{equation}\label{ogroup}
G = \{ 1, (-1)^{F_L}R_{45}\Omega,(-1)^{F_L} R_{67}\Omega,R_{4567} \},
\end{equation}
where $R$ reflects the coordinates indicated and $\Omega$ is the worldsheet
parity.
The splitting of the D6-branes into half-D6-branes discussed in
\cite{karch} becomes obvious after T-duality. Indeed, it follows easily from
the above formulas that the location of the upper half 6-branes,
$X_4=X_5=0, X_7>0$ in the type IIA configuration maps to the locus $z_2=0$ in
IIB. Similarly, the lower halfs of the 6-branes, $X_4=X_5=0, X_7<0$,
map to $z_1=0$. Thus the upper halfs of D6-branes map to {\it whole}
D7-branes located at $z_2=0$, while the lower halfs map to {\it whole}
D7-branes at $z_1=0$.
To specify the theory on the 7-branes completely we need to make a
consistent choice for the action of the orientifolds on the Chan-Paton factors
of the 7-7, 7-7$'$, and 7$'$-7$'$ strings. There are at least two such choices.
One gives rise to an $SO(8)\times SO(8)$ gauge symmetry \cite{bz}, and
classically the
other yields a $U(4)\times U(4)$ gauge group on the 7-branes \cite{sen1,sen2},
which is broken to $SU(4)\times SU(4)$ by one-loop effects~\cite{berkoozetal}.
The second case is related to the Gimon-Polchinski \cite{gp} orientifold via
T-duality. We will be mainly interested in the second orientifold, which we
will refer to as the
Sen model. Both of these orientifolds were constructed as compact models
with a total of four orientifolds and sixteen physical 7-branes of each kind.
The 7-brane gauge groups listed here are the parts of the total 7-brane
group that are visible to a 3-brane probe near one of the intersections.
The theory on a 3-brane probe in the Sen model background was analyzed
in \cite{ah1}. The gauge group, matter content, and the superpotential are in
complete agreement with the theory we discussed in section \ref{conformal}.
Thus we conclude that the IIA configuration with all 6-branes on top of the
NS5-brane is T-dual to a local piece of the Sen model \cite{sen1,sen2}.
As in the IIA description the flat directions of the field theory correspond
to motions of the 3-branes in the 7-brane background. Moving the 3-branes
off the intersection point along either of the O7-planes corresponds to giving
en expectation value to one of the bifundamentals ${\cal Q},\tilde{\cal Q}$,
and moving the 3-branes off both orientifolds gives an expectation value to
both ${\cal Q}$ and $\tilde{\cal Q}$. Separating the
3-branes in the direction which the 7- and 7$'$-branes share corresponds to
giving expectation values to the antisymmetric tensors $A_1,A_2$.
It is instructive to study the deformations
of the Sen model and compare these to the deformations of the
corresponding IIA construction. The IIA construction has the advantage that
all deformations of the background correspond to moving the 6-branes or the NS5-branes. In
the IIB picture only some of the deformations are geometric, others correspond
to Wilson lines. Once the map between IIA and IIB deformations is established,
we can also find the IIB description of the second (non-conformal) IIA configuration
discussed in section \ref{iic}.
Sen \cite{sen1,sen2} has studied the deformations of the compact model in
great detail. In the compact case the field theory on the 7-branes turns out
to be a $(1,0)$ theory in six dimensions. Since our IIB configuration is non-compact,
we cannot simply use Sen's results.
In fact, in our case the theory on the 7-branes is not even six-dimensional, instead
it is an eight-dimensional theory with six-dimensional
impurities. Such theories have been discussed previously~\cite{kap2,kapdn}.
Before we launch into an analysis of the impurity theory we need to discuss
the matter content of the 7-brane theory. A single O7$^-$-plane with four
coincident 7-branes gives rise to an ${\cal N}=1$ $SO(8)$ theory in eight dimensions.
The bosonic degrees of freedom in the eight-dimensional vector multiplet consist of a
vector field and a complex scalar, both in the adjoint of the gauge group.
The second O7$^-$-plane in our configuration breaks half of the supersymmetries
and imposes projections on fields in the vector multiplet. With the
projection matrices for the Sen model \cite{gp,sen1}, the surviving constant modes of
the fields are a vector and a complex scalar in the ${\bf 6}+
\bar{\bf 6}$. These fields account for the 7-7 strings and there are similar fields on
the 7$'$-branes from 7$'$-7$'$ strings. The 7-7$'$ strings are localized at the intersection
of 7- and 7$'$-branes. They yield a single hypermultiplet of the six-dimensional $(1,0)$
theory on the intersection, which transforms as a $({\bf 4},{\bf 4})$ under the
(classical) $U(4)_7\times U(4)_{7'}$ gauge group.
\subsection{The seven-brane impurity theory}
In this section we analyze the supersymmetric vacua of the impurity theory on
the 7-branes and compare them with the vacua of the T-dual IIA configuration.
We expect the vacuum field configurations to be translationally invariant in
the six directions common to the 7- and 7$'$-branes. Focusing now on the
7-branes,
we see that we can capture the physics by studying the dependence of the
7-brane fields on the remaining two directions transverse to the 7$'$-branes.
The 7$'$-branes and the O7$'$-plane intersect this two-dimensional plane
in a point.
To set up the impurity theory we use a complex affine coordinate $z$ on the plane
and define $A_{\bar z} = \frac{1}{2}(A_1+iA_2)$, where $A_i$ are the two
components of the $SO(8)$ gauge field living on the 7-branes.
The 7-brane theory also contains a complex scalar, $\Phi$, in the adjoint of
$SO(8)$ that describes the transverse fluctuations of the 7-branes.
The bifundamental $({\cal M},\tilde{\cal M})$ from the 7-7$'$ strings is localized
at the point $z=0$. A very similar theory (without orientifold projections)
was described in \cite{kap2}. The moduli space of the
impurity theory is given by the solution of the equations
\begin{eqnarray}\label{imp}
&F_{z\bar{z}}-[\Phi,\Phi^\dagger] =
\delta(z)\left({\cal MM}^\dagger
-\tilde{\cal M}^\dagger\tilde{\cal M}\right)&\nonumber \\
&\bar{D}\Phi=-\delta(z) {\cal M}\tilde{\cal M},&
\end{eqnarray}
where $F_{z\bar{z}}= \partial A_{\bar z} - {\bar \partial}A_z-[A_z,A_{\bar z}]$
and ${\bar D}={\bar\partial}-A_{\bar z}$.
These equations are known as Hitchin equations with sources. They are
analogous to the $D$ and $F$ flatness conditions in ordinary supersymmetric field
theories. A similar set of equations describes the impurity theory on the
7$'$-branes.
To make contact with the notation in \cite{sen1,sen2} we write all 7-brane
fields as antisymmetric $8\times 8$ matrices with certain constraints on the entries.
This reflects the origin of the fields in the impurity theory. Without the
O7$'$-plane, both $A_{\bar z}$ and $\Phi$ would transform in the adjoint of
$SO(8)$. Orientifolding with O7$'$ puts additional constraints on these fields
\begin{eqnarray}\label{constr}
\Phi(z) &=& P\Phi^T(-z)P^{-1} \nonumber \\
A(z) &=& PA^T(-z)P^{-1},
\end{eqnarray}
where
\begin{equation}
P = \left( \begin{array}{cc} P_4 & 0 \\ 0 & -P_4 \end{array} \right), \qquad
P_4 = \left( \begin{array}{cc} 0 & {\bf 1}_{2\times 2} \\
-{\bf 1}_{2\times 2} & 0 \end{array} \right).
\end{equation}
Orientifolding also breaks the gauge group from $SO(8)$ down to the group of all
continuous $SO(8)$-valued functions satisfying $g(z)=P g(-z) P^{-1}$. In particular,
at $z=0$ the gauge group reduces to $U(4)$.
The orientifold projections allow the bifundamentals to be arbitrary complex
$8\times 8$ matrices that commute with $P$ \cite{gp}.
The impurity equations are consistent if the products of the bifundamentals
on the r.h.s.~of \eq{imp} are antisymmetrized in the gauge indices.
We need to find all, possibly $z$-dependent, field configurations that satisfy
the impurity equations, \eq{imp}, modulo gauge transformations. To this end we make
the following ansatz
\begin{equation}
A_{\bar z} = \frac{T}{z}, \qquad \Phi(z) = \Phi_0 + \frac{\Phi_s}{z}.
\end{equation}
Here $T, \Phi_0$ and $\Phi_s$ are constant antisymmetric $8\times 8$ matrices.
Imposing the constraints, \eq{constr}, determines that $\Phi_0$ transforms in
the ${\bf 6}+\bar{\bf 6}$ of $U(4)$ while $T$ and $\Phi_s$ transform as
adjoints. The background gauge field, $A_{\bar z}$, can be interpreted as a
flat connection that gives rise to a Wilson line around the intersection point at
$z=0$. The constant part of the scalar field, $\Phi_0$, corresponds to the asymptotic
(i.e. $z\to\infty$) positions of the 7-branes in the directions transverse to the
O7-plane, while the singular part, $\Phi_s$, parametrizes a deformation of the shape of
the 7-branes.
The moduli space of the impurity equations, \eq{imp}, has several branches
with rather different physics. The simplest situation arises if all
bifundamental expectation values and all singular parts of $A_{\bar z}$ and $\Phi$
vanish. In that case \eq{imp} reduces to the condition
\begin{equation}
[\Phi_0,\Phi_0^\dagger] = 0,
\end{equation}
which is solved by
\begin{equation}\label{six}
\Phi_0 = \left( \begin{array}{cc} 0 & \phi \\ -\phi & 0 \end{array}\right),
\qquad
\phi = {\rm diag}(\phi_1,\phi_2,\phi_1,\phi_2).
\end{equation}
As in Ref.~\cite{sen1}, the two complex parameters, $\phi_{1,2}$, parametrize
the transverse position of two pairs of 7-branes. We discuss the corresponding
IIA deformation in the next section. For the remainder of this section we
set $\Phi_0=0$.
The impurity equations, \eq{imp}, become inhomogeneous once we turn on an
expectation value for the bifundamental fields. Since
${\bar\partial}(1/z) \sim \delta(z)$, and the r.h.s.~of \eq{imp} is
proportional to $\delta(z)$, the singular fields above have the
right form to satisfy the impurity equations with nonzero bifundamental
expectation values.
The most generic expectation value of the bifundamentals for which the
impurity equations have solutions reads
\begin{eqnarray}
&{\cal M} = \left(\begin{array}{cc} M_1 & 0 \\ 0 & M_2 \end{array} \right),&\\
& M_1 = \left( \begin{array}{cccc}
m_1 & 0 & -im_1 & 0 \\
0 & m_2 & 0 & -im_2 \\
im_1 & 0 & m_1 & 0 \\
0 & im_2 & 0 & m_2 \end{array} \right), \qquad
M_2 = \left( \begin{array}{cccc}
m_3 & 0 & im_3 & 0 \\
0 & m_4 & 0 & im_4 \\
-im_3 & 0 & m_3 & 0 \\
0 & -im_4 & 0 & m_4 \end{array} \right),&\nonumber
\end{eqnarray}
and an expectation value of the same form, but with $m_i$ replaced by $\tilde{m}_i$,
for $\tilde{\cal M}$.
The impurity equations determine the expectation values of the other fields
in terms of ${\cal M}$ and $\tilde{\cal M}$. The residue of $\Phi$
is given by
\begin{equation}\label{x45}
\Phi_s = {\rm diag}(\Phi_1,\Phi_2),
\end{equation}
where
\begin{equation}
\Phi_1 = \left( \begin{array}{cccc}
0 & 0 & -\phi_1 & 0 \\
0 & 0 & 0 & -\phi_2 \\
\phi_1 & 0 & 0 & 0 \\
0 & \phi_2 & 0 & 0 \end{array} \right), \qquad
\Phi_2 = \Phi_1(\phi_1\to-\phi_3,\phi_2\to -\phi_4),
\end{equation}
with $\phi_i \sim m_i\tilde{m}_i$. The matrix $T$ in the gauge connection has
the same structure as $\Phi_s$, except that $\phi_i$ is replaced by
$t_i \sim |m_i|^2 - |\tilde{m}_i|^2$.
Before discussing this general solution, we will focus on two special cases.
If we set $m_i = \tilde{m}_i$, the r.h.s.~of the first impurity equation
vanishes and only the residue of $\Phi$ is turned on.
This expectation value of the bifundamentals
breaks the $U(4)\times U(4)$ gauge group
to a diagonal
subgroup. If all $m_i$ are equal this subgroup is $U(4)_D$, and for generic
values of $m_i$ we find $U(1)^4$. Since the
7-brane group is broken to a diagonal subgroup, the impurity theory,
\eq{imp}, on the 7-branes and the corresponding impurity theory on the 7$'$-branes
contain the same information. Therefore it is sufficient to consider only the
7-brane theory. The field $\Phi(z)$ describes the
shape of the 7-branes. For large $z$ the 7-branes asymptote to the O7-plane
as in the unperturbed case, while they approach the O7$'$-plane for small $z$.
Thus we conclude that turning on this bifundamental expectation value deforms
pairs of intersecting 7- and 7$'$-branes into a single smooth 7-brane that
interpolates between the 7- and 7$'$-branes.
This result agrees with the F-theory analysis in \cite{sen2}, where this
behavior was interpreted as fusing the 7- and 7$'$-branes together.
There are also solutions of the impurity equations with non-zero gauge
connection and $\Phi_s=0$. We find one such solution if we set
$m_1 = m_2=\tilde{m}_3=\tilde{m}_4$,
and all other components of the bifundamentals vanish. For this choice the
r.h.s.~of the second equation in \eq{imp} vanishes, which implies $\Phi_s=0$,
and $t_1=t_2 \sim |m_1|^2$, $t_3=t_4 \sim -|m_1|^2$.
This bifundamental expectation value breaks the $U(4)_7\times U(4)_{7'}$
7-brane gauge group to a diagonally embedded $U(2)\times U(2)$.
Note that this deformation is purely non-geometric. Since $\Phi(z)=0$, the
7-branes have the same shape as in the case without any bifundamental
expectation values.
It is now a simple matter to identify these two singular solutions with the
corresponding deformations in the IIA construction. The first solution with
$T=0$, $\Phi_s \neq 0$
corresponds to moving the 6-branes off the NS5-brane in the $X_{4,5}$
direction. If none of the 6-branes coincide, the $U(4)\times U(4)$ gauge
symmetry on the 6-branes is broken to $U(1)^4$. This is in complete agreement
with the impurity analysis. Note
that a deformation that corresponds to fusing 7 and 7$'$-branes together in
the IIB description maps into a simple brane motion in the IIA construction,
which involves reconnecting the upper and lower halfs of the 6-branes.
The second singular solution with $T\neq 0$, $\Phi_s = 0$
also corresponds to a simple brane
motion in the IIA description. We identify turning on $m_1$
with the motion of two pairs of 6-branes in the $X_6$ direction.
The classical gauge group on the 6-branes is $U(2)\times U(2)$ as expected
from the
IIB analysis. This brane motion also requires that we reconnect the upper and
lower halfs of the 6-branes, so that the resulting 6-brane group is a diagonal
subgroup of the original $U(4)\times U(4)$ gauge symmetry. This is in
perfect agreement with the analysis of the 7-brane impurity theory.
It is straightforward to discuss more general choices for the bifundamental
expectation values. The bifundamental expectation values are parametrized
by eight complex numbers, $m_i$ and $\tilde{m}_i$, which determine the
matrices $T$ and $\Phi_s$ completely. The four parameters in $T$ map into
the $X_6$ position of the 6-branes in the IIA description and the entries
in $\Phi_s$ correspond to the $X_{4,5}$ positions. Thus we find complete
agreement between the brane motions in the IIA description and the moduli
corresponding to singular fields in the impurity theory.
\subsection{A supersymmetric IIA configuration with curving six-branes}
The deformations we discussed so far are rather complicated in the IIB
picture and correspond to simple brane motions in the IIA description.
In fact, all simple brane motions in the IIA description are accounted for.
However, there is
a very simple brane motion in IIB, namely the constant solution of the impurity equations
given in \eq{six}, that should have a counterpart in the IIA description.
Since this deformation corresponds to moving pairs of 7-branes off the
orientifold, we can find an explicit equation describing the position of these
branes. In terms of the coordinates in \eq{z12} this equation reads
$z_2 = const$. Starting from this expression
we can reverse the coordinate transformations that took us from the Taub-NUT
space to the flat coordinates on ${\bf C}^2/{\bf Z}_2$. This provides an
expression for the world volume of the 7-brane in the Taub-NUT coordinates.
Since the 7-branes wrap the fiber of the Taub-NUT and the fiber T-dualizes to
the compact $X_6$ direction, it is straightforward to find the equation for the
worldvolume of the corresponding 6-brane. The result is $X_4^2-c X_7-c^2/4=0$, i.e., a
parabola in the $X_4-X_7$ plane. Fig.~\ref{fig2} shows the IIA
configuration which is T-dual to the following IIB situation: all 7$'$-branes are
coincident with the O7$'$-plane, and one pair of 7-branes is displaced from
the O7-plane.
\begin{figure}
\centerline{\epsfxsize = 5truecm \epsfbox{fig2.eps} }
\tighten{
\caption{ Type IIA configuration for non-zero expectation value of the {\bf 6}.
The dot represents the NS5-brane, the thick line corresponds to four half
6-branes, the thin line corresponds to two half 6-branes and the curving line
is another 6-brane. }
\label{fig2}
}
\end{figure}
From this picture one can see that turning on the constant complex scalar on the 7-brane
corresponds to fusing two upper halfs of the 6-branes together and moving them off the
NS5-brane as shown in the figure. On the IIB side it is obvious that
this deformation preserves all supersymmetries. This is somewhat harder to see on the IIA side.
Presumably the $H$-field produced by the NS5-brane stabilizes the curved worldvolume of the
D6-brane.
The effect of this deformation on the probe theory is what we expect from
the IIB picture. There we move two 7-branes away from the 3-branes sitting at
the intersection point of the orientifold planes. This gives a mass to half of
the fundamentals from 7-3
strings. In the IIA picture the deformation accomplishes the same. In the
IIB picture moving the 3-branes along the O7$'$-plane and transverse to the
O7-plane corresponds to
giving the bifundamental field ${\cal Q}$ in the probe theory an expectation value
\cite{ah1}. Thus it is possible to move the 3-branes away from the intersection of the
orientifolds towards the intersection of the pair of 7-branes with the O7$'$-plane by giving
an expectation value to one of the bifundamentals. This is also reflected in the
IIA description. We can move the 4-branes in the negative $X_7$ direction by giving an
expectation value to one of the bifundamentals (see section \ref{iib}). This moves the
4-branes off the NS5-brane and towards the intersection of the lower half-6-branes with
the curving 6-brane.
In the IIB description moving a pair of 7-branes away from the O7-plane breaks the
7-brane gauge group from $SU(4)$ down to $SU(2)\times SU(2)$~\cite{sen1,sen2}.
Moving all four 7-branes together breaks $SU(4)$ down to $Sp(4)$. This implies that
the unbroken gauge group on a single curving 6-brane should be $SU(2)$,
while for two coincident curving branes it should be enhanced to $Sp(4)$.
It is not at all clear how to see this from the IIA description.
\subsection{Comparison with F-theory}
Sen argued \cite{sen1,sen2} that the T-dual version of the GP model
\cite{gp} is related to an F-theory compactification with certain fluxes through
collapsed 2-cycles. The naive candidate for such an F-theory compactification
would be a pair of intersecting $D_4$ singularities. However, this cannot be
directly related to the GP orientifold, since it would give rise to an
$SO(8)\times SO(8)$ gauge symmetry and contain tensionless strings, while the GP model
has $SU(4)\times SU(4)$ symmetry and no tensionless strings. The difference is due to
NS (and possibly RR) 2-form fluxes through the collapsed 2-cycle at the intersection of the
two $D_4$ singularities. These fluxes give a mass to 3-branes wrapping this
cycle, thereby preventing the appearance of tensionless strings. These fluxes
are not quantized \cite{sen1,sen2}, so we should be able to identify moduli
in our IIA description that correspond to turning them off. The NS flux is
conventionally identified with the position of the NS5-branes on the $X_6$
circle and the RR flux parametrizes the location of the NS5-branes on the
M-theory circle. From the IIB point of view, they are both part of a massless
hypermultiplet living at the intersection of the $D_4$ singularities. In order to turn
off the NS flux, we move the NS5-brane and its image as well as all D6-branes to
coincide with one of the O6-planes.
This configuration has an $SO(8)\times SO(8)$ gauge symmetry from
the eight upper and eight lower halfs of the 6-branes, as well as tensionless strings from
the NS5-brane coincident with its image \cite{hz}.
In addition to the hypermultiplet that corresponds to moving the NS5-brane
off the orientifold in the $X_4,X_5,X_6,X_{10}$ there is now a tensor multiplet whose
scalar expectation value corresponds to separating the two NS5-branes in the $X_7$ direction.
All this agrees with the expectations from F-theory.
\section{The large $N$ limit}
\label{sugra}
When the number of D3-branes, $N$, is large there is a dual description of
${\cal N}=1$ superconformal theory on the D3-branes in terms of a supergravity
on $AdS_5\times X$,
where $X$ is an Einstein manifold (or orbifold)~\cite{m1}.
This dual description is valid when the t'Hooft gauge coupling, $g_{YM}^2 N$,
is large. In this section we will show how the AdS/CFT correspondence works
for the conformal gauge theory with $SU(4)\times SU(4)$ flavor symmetry discussed in
section \ref{iib}, and provide evidence that this theory is finite. We will also provide
a partial analysis of the non-conformal theory of section \ref{iic} in the large $N$ limit
and argue that supergravity suggests a definite R-charge assignment for all the fields
in the infrared.
\subsection{The conformal case}
In the conformal case, $X$ is an orientifold of ${\bf S}^5$. As explained in
the previous section, the IIA configuration with $SU(4)\times SU(4)$ gauge
symmetry on the 6-branes is T-dual to a local piece of the Sen model. At the
$SU(4)\times SU(4)$ point, the Sen model is a perturbative type IIB
orientifold with constant string coupling, $\tau$ \cite{sen1,sen2}. Thus the
near-horizon geometry of the 3-branes is obtained by orientifolding $AdS_5\times {\bf S}^5$.
Similar theories were analyzed in \cite{f1,f2,kap}.
Let us denote the orientifolded five-sphere by ${\tilde{\bf S}}^5$.
The metric on ${\tilde{\bf S}}^5$ is the angular part of
\begin{equation}
ds^2 = |dz_1|^2 + |dz_2|^2 + |dw|^2,
\end{equation}
where $w=X_8+iX_9$ and the variables $z_1,z_2$ are subject to the
identifications $z_i \to -z_i$. A $U(1)^3$ subgroup of the $SO(6)$ isometry
group of ${\bf S}^5$ commutes with these identifications. It is convenient
to take the generators that rotate $z_1$, $z_2$,and $w$ separately as a basis
in the Lie algebra of $U(1)^3$. Explicitly, the metric on ${\tilde{\bf S}}^5$
can be written as
\begin{equation}
ds^2_{{\tilde{\bf S}}^5} = d\theta_1^2+\sin^2(\theta_1)d\phi_1^2 + \cos^2(\theta_1)
\left( d\theta_2^2+\sin^2(\theta_2)d\phi_2^2+\cos^2(\theta_2)d\phi_3^3\right),
\end{equation}
where $\phi_{1,2} \in [0,\pi]$, $\phi_3 \in [0,2\pi]$, and
$\theta_{1,2} \in [0,\pi]$. The three angles $\phi_i$ parametrize rotations
in the $z_{1,2}$ and $w$ planes respectively. The periodicity of $\phi_{1,2}$
reflects the identifications on $z_{1,2}$. Since this periodicity of $\phi_{1,2}$
is the only thing which distinguishes ${\tilde{\bf S}}^5$ from ${\bf S}^5$, the eigenvalues
of the scalar Laplacian on the former can be obtained from those on the latter.
The eigenvalue of the scalar Laplacian on ${\bf S}^5$ is $k(k+4)$, where $k=0,1,\ldots$.
In terms of the angular momenta, $m_i$, associated with the angles $\phi_i$,
we have
$k = |m_1|+|m_2|+|m_3|+2l_1+2l_2$, where $l_i$ are non-negative integers.
The orientifold projection on the bulk supergravity states amounts to
keeping modes with even $m_1$ and $m_2$.
In the ${\cal N}=4$ case, the supergravity states with lowest mass squared
come from the KK reduction of $h^a_a$, the dilaton mode of the
${\bf S}^5$. The AdS masses of these states are given by $m^2 = k(k-4)$ \cite{romans},
where $k$ is given above. According to \cite{w1}, the AdS mass of a KK state is
related to the dimension of the corresponding boundary operator by
$\Delta(\Delta-4)=m^2$, which implies $\Delta=k$ for this tower of KK modes.
The decomposition of the other supergravity fields yield towers of KK states
for which $\Delta = k+n$, where $n$ is a positive integer \cite{w1}.
We will see below that only for $n=0$ the KK states couple to chiral primary operators.
Therefore we will restrict our analysis to the KK modes from the
decomposition of $h^a_a$.
The simplest way to identify chiral primaries is to find all states for which
$\Delta=\frac{3}{2}R$, where $R$ is the R-charge which is part of the superconformal algebra.
The R-current is a certain linear combination of the three $U(1)$ currents.
To find this linear combination we first need to determine which supercharges
survive the orientifold projection. The orientifold group, ${\bf Z}_2\times
{\bf Z}_2$, is generated by $\gamma_1=R_{z_2}\Omega (-1)^{F_L}$ and
$\gamma_2=R_{z_1z_2}$.
Orientifolding by the first generator breaks $SO(6)$ down to $SU(2)_L\times SU(2)_R
\times U(1)_N$ where $U(1)_N$ acts on $z_2$ while $SU(2)_L\times SU(2)_R$ acts
on $z_1,w$. The surviving supercharges $(Q_+,Q_-)$ transform as $({\bf 1},{\bf 2})_1$
with respect
to this group. Orientifolding by $\gamma_2$ breaks $SU(2)_L\times SU(2)_R$ down to
$U(1)_L\times U(1)_R$. We will denote the sum of the $U(1)_L$ and $U(1)_R$
charges
by $U(1)_2$, the difference by $U(1)_3$, and refer to $U(1)_N$ as $U(1)_1$.
The charges of $z_2, z_1,$ and $w$ under these three $U(1)$'s are given by
$(2,0,0),(0,2,0),$ and $(0,0,2)$, respectively. The supercharge $Q_+$ which survives the
second orientifolding has $U(1)$ charges $(1,1,-1)$. It follows that the R-charge which
is in the same superconformal multiplet as the stress-energy tensor is
$\frac{1}{3}(2m_1+2m_2-2m_3)$. Here $2m_1$ is the $U(1)_1$ charge, $2m_2$ is the
$U(1)_2$ charge, and $2m_3$ is the $U(1)_3$ charge.
The normalization is chosen so that $Q_+$ has R-charge $1$.
It follows that any KK mode with
$l_1=l_2=0$, $m_1,m_2 \ge 0$ and $m_3 \le 0$ should couple to a chiral
primary operator in the boundary field theory.
We discussed the identification of geometric motions of 3-branes with flat
directions in the 3-brane field theory in the previous section (see also
\cite{ah1}). This allows us to
determine the $U(1)$ charges of the fields $A_1,A_2,{\cal Q},\tilde{\cal Q}$.
The field theory superpotential then fixes the R-charges of the fundamentals
$q,{\tilde q},p,{\tilde p}$. The results are summarized in the table below.
\begin{equation}\label{table}
\begin{array}{c|ccc}
& U(1)_1 & U(1)_2 & U(1)_3 \\ \hline
A_{1,2} & 0 & 0 & -2 \\
{\cal Q} & 2 & 0 & 0 \\
\tilde{\cal Q} & 0 & 2 & 0 \\
q,p & 0 & 1 & -1 \\
\tilde{q},\tilde{p} & 1 & 0 & -1
\end{array}
\end{equation}
With these charge assignments in hand it is now a simple matter to match
the bulk KK modes and the chiral primary operators in the field theory. Let us
give some examples.
The supergravity spectrum contains a singleton chiral primary with
$U(1)_3$ charge $-2$ and $\Delta =k=1$.
This state corresponds to a chiral primary ${\rm Tr} (A_1 J_1) + {\rm Tr} (A_2 J_2)$ in the field
theory.\footnote{The antisymmetric representation of $Sp(N)$ is reducible and
contains a singlet.} Since $\Delta=1$, this is a free field. For $\Delta = 2$
there are three chiral primary states with geometric $U(1)$ charges $(4,0,0)$, $(0,4,0)$,
and $(0,0,-4)$. We identify them with ${\rm Tr}{\cal Q}^T J_1 {\cal Q} J_2$,
${\rm Tr}\tilde{\cal Q}^T J_2 \tilde{\cal Q} J_1$, and ${\rm Tr} (A_1 J_1)^2 + {\rm Tr} (A_2 J_2)^2$.
The chiral primary operators with $\Delta = 3$ are
${\rm Tr}[{\cal Q} A_2 {\cal Q}^T J_1+{\cal Q}^T J_1 A_1 J_1 {\cal Q} J_2]$,
${\rm Tr}[\tilde{\cal Q} A_1 \tilde{\cal Q}^T J_2+
\tilde{\cal Q}^T J_2 A_2 J_2 \tilde{\cal Q} J_1]$, and
${\rm Tr} (A_1 J_1)^3 + {\rm Tr} (A_2 J_2)^3$. They correspond to the KK states with
charges $(4,0,-2)$, $(0,4,-2)$, and $(0,0,-6)$ respectively.
The field theory also contains operators that carry charges under the
7-brane gauge groups. It was pointed out in \cite{f2} that these operators
couple to the AdS modes coming from the KK reduction of the 7-brane fields.
Our configuration includes an O7-plane with four coincident D7-branes
wrapping an ${\bf S}^3$ defined by $|z_1|^2+|w|^2=const$,
and similarly an O7$'$-plane
with four D7$'$-branes wrapped on $|z_2|^2+|w|^2=const$.
The two 3-spheres intersect over a circle.
We can focus on the KK modes from the first ${\bf S}^3$. These modes couple
to operators that are charged under the $SU(4)_7$ subgroup of the
$SU(4)_7\times SU(4)_{7'}$
global symmetry group of the probe theory. The modes living on the other
${\bf S}^3$ couple to similar operators in the field theory that transform
under $SU(4)_{7'}$.
The KK reduction of the theory on an O7-plane with four coincident 7-branes
was discussed in \cite{f2}. In that case there were twice as many supersymmetries as in
ours. The simplest way to compute the KK spectrum in our case is to
use the results of \cite{f2} and impose the additional projection from the
O7$'$-plane.
Ref.~\cite{f2} contains a detailed discussion of the 7-brane states and
their multiplet structure. The lowest component of the multiplet is a
real field in the $({\bf k}, {\bf k+2})_0$ representation of
$SU(2)_L\times SU(2)_R\times U(1)_N$, where $k=1,2,\ldots$.
This mode comes from KK reduction of the components of the 7-brane gauge
field along the ${\bf S}^3$,
\begin{equation}
A_a = \sum_k a_k Y_a^k,
\end{equation}
where $Y_a^k$ is the $k$-th vector spherical harmonic on ${\bf S}^3$. These modes
couple to operators of dimension $\Delta = k+1$ in the boundary field theory.
For simplicity we will only consider operators with $\Delta = 2,3$.
The state with $\Delta=2$ transforms
in the $({\bf 1},{\bf 3})_0$ and decomposes into
modes with $U(1)^3$ quantum numbers $(0,0,0)$ and $(0,\pm 2, \mp 2)$.
The $(0,0,0)$ mode has no $U(1)_R$ charge and does not correspond to a chiral primary.
The states with $U(1)^3$ charges $(0,2,-2)$ and $(0,-2,2)$ are complex
conjugates of each other, so it is sufficient to consider only one of them,
e.g., the first. It has R-charge $4/3$ and is, therefore, a chiral primary.
This state starts out in the adjoint of the $SO(8)_7$ gauge group on the
7-brane. Since it has $m_2=1$, it is odd under the additional orientifold
projection $\gamma_2$. This projection breaks $SO(8)_7$ down to $SU(4)_7$. As
explained in~\cite{sen1,sen2}, states in the adjoint of $SO(8)_7$ which are
odd under $\gamma_1$ yield ${\bf 6}+\bar{\bf 6}$ of $SU(4)_7$, while even states
give adjoints of $SU(4)_7$. It follows that the
$(0,2,-2)$ state yields one complex state in ${\bf 6}$ and one complex state
in $\bar{\bf 6}$. These KK states correspond to operators $qJ_1q$ and $pJ_2p$, which
transform in the $\bf 6$ and $\bar{\bf 6}$ of the 7-brane group respectively.
The $\Delta=3$ mode is in the $({\bf 2},{\bf 4})_0$ representation and decomposes
into even modes with $U(1)^3$ charges $(0,0,-2)$ and $(0,4,-2)$ and their complex
conjugates, as well as odd modes with $U(1)^3$ charges $(0,2,0)$ and $(0,2,-4)$ and
their complex conjugates.
The even $(0,0,-2)$ mode and the odd $(0,2,0)$ mode do not couple to chiral primary operators,
because the R-charge does not match the dimension. The even $(0,4,-2)$
mode couples to a chiral primary operator in the adjoint of $SU(4)$ which we
identify as $pJ_2\tilde{\cal Q}J_1q$. The odd $(0,2,-4)$ mode couples to a
chiral primary in the ${\bf 6}+\bar{\bf 6}$. The corresponding operators are
given by $qA_1q$ and $pJ_2A_2J_2p$.
Other scalars on AdS come from the decomposition of the complex scalar
field on the 7-branes. These KK modes are in the $({\bf k},{\bf k})_2$
representation of $SU(2)_L\times SU(2)_R\times U(1)_N$ and couple to operators of
dimension $k+2$~\cite{f2}. It is straightforward to decompose and project these modes as we
did for the KK modes of the vector field. The $\Delta=3$ case is especially
simple, since this mode carries only $U(1)_1$ charge. Since the R-charge
and the dimension do not satisfy $\Delta=\frac{3}{2} R$, this KK mode does not
couple to a chiral primary operator. The same is true for the higher KK modes of
the complex scalar field.
Finally, there are also states living on the intersection of the 7-branes and 7$'$-branes
which is an ${\bf S}^1$ embedded in ${\bf S}^5$. The KK reduction of these states is
straightforward, and we will not discuss it.
In the above analysis we have focused on chiral primaries. It is also interesting
to ask whether non-chiral states match between field theory and supergravity.
Some of the non-chiral scalars we have seen, namely the ones coming from the
reduction of complex scalars living on the 7-branes, are descendants of the
chiral primaries and therefore match automatically. On the other hand, the
non-chiral scalars which come from the KK reduction of the gauge field on the 7-branes
are primary. One may ask whether the superconformal multiplet they live in is
long or short.
To answer this question we need to recall some
facts about unitary representations of the ${\cal N}=1$ superconformal algebra~\cite{conf}.
For our purposes it is sufficient to consider multiplets whose primary states have zero spin.
Let the $R$ and $\Delta$ be the R-charge and the dimension of the primary. Unitarity
puts restrictions on which values of $R$ and $\Delta$ may occur; the allowed possibilities
are
(i) $\Delta=R=0$ (the trivial representation),
(ii) $\Delta=\frac{3}{2}|R|$ (chiral and anti-chiral representations),
(iii) $\Delta\geq\frac{3}{2}|R|+2$.
Representations of type (iii) with $\Delta>\frac{3}{2}|R|+2$ contain no null states
and therefore are termed long multiplets. Chiral and anti-chiral representations contain
null states at level one, i.e., their primaries are annihilated by half of the supercharges.
These representations are called short.
Representations of type (iii) which saturate the inequality are also short; the null states
occur at level two. A well-known example of a short multiplet is a linear multiplet which
contains a conserved current. It corresponds to the case $R=0,\Delta=2$.
One can check that all non-chiral primaries coming from the reduction of the gauge field
on the 7-branes satisfy $\Delta=\frac{3}{2}|R|+2$ and therefore are in short multiplets
of type (iii).
In particular, the $(0,0,0)$ mode with $\Delta=2$ we have found above
is in fact the lowest component of a linear multiplet. It couples to a field theory operator
$q^\dagger q - pp^\dagger$ in the adjoint of $SU(4)_7$. The corresponding current
is simply the $SU(4)_7$ flavor current.
The matching of non-chiral primaries with $\Delta=3$ is a bit more involved.
The $(0,2,0)$ mode transforms in ${\bf 6}+\bar{\bf 6}$ of $SU(4)_7$. Its
field theory
counterparts are $h_1 p^\dagger\tilde{\cal Q}J_1 q + h_2 q J_1 A_1^\dagger J_1 q$ and
$h_1 pJ_2\tilde{\cal Q} q^\dagger+h_2 p A_2^\dagger p$, where the flavor indices
are antisymmetrized. The $U(1)^3$ charges of these operators match those of the
$(0,2,0)$ mode. To show that these operators live in short multiplets,
i.e., are annihilated by $\bar D^2$, one needs to use the classical equations
of motion. The manipulations one has to go through are very similar to those
in~\cite{berkooz}, and are subject to the same caveats. The use of the classical equations of
motion is presumably justified in the weakly coupled regime where $g^2_{YM} N$ is small. The
supergravity analysis indicates that the operators in question belong to short
multiplets even for large $g^2_{YM} N$. An even more interesting
situation arises when one tries to match the non-chiral primary with $U(1)^3$ charges
$(0,0,-2)$ and $\Delta=3$. This mode lives in the adjoint of $SU(4)_7$. We
claim that it corresponds to the field theory operator
$h_1 q A_1 J_1q^\dagger-h_1 p^\dagger A_2 J_2p-h_2 q \tilde{\cal Q}^\dagger p$.
Evaluating the ${\bar D}^2$ descendant of this operator using the
classical equations of motion, one finds that it does not vanish. Instead, the
descendant has the form $(q\tilde{q})(\tilde{p}p)$, i.e., it factorizes into a
product of two gauge-invariant operators and is therefore subleading at large
$N$. It follows that this field theory operator lives in a long
multiplet for finite $N$, but is ``close'' to being in a short multiplet
in the sense that its dimension approaches the unitarity bound as $N\to\infty$.
On the supergravity side this means that the $(0,0,-2)$ one-particle state is in a short
multiplet only for $N=\infty$. For finite $N$ the multiplet absorbs another short
multiplet made of two-particle states and becomes long.
This concludes our analysis of the AdS/CFT correspondence for the Sen model.
There is complete agreement between the spectrum of primary operators
in the field theory and the scalar Kaluza-Klein states on AdS as required
by the AdS/CFT correspondence~\cite{m1,gubs,w1}. The charge assignments in
Table~\ref{table}
together with the formula $R=\frac{1}{3}(2m_1+2m_2-2m_3)$ imply that all
chiral fields have canonical dimensions in the infrared. This is the most
natural assumption for a theory with
vanishing beta function, but as we pointed out in the introduction there is no
field theory proof of this. The supergravity computation is only valid for
large $N$ and large $g^2_{YM} N$. However, given that for $g^2_{YM}\ll 1$ and
$N$ of
order $1$ the dimensions are also canonical, it appears likely that the theory is
finite for all $N$.
\subsection{The non-conformal case}
Next we discuss the deformed ${\cal N}=1$ theory which flows to a line of conformal
fixed points in the infrared (section \ref{iic}). We have already pointed out that
although the Wilsonian gauge coupling in this theory depends on the scale,
the low-energy effective gauge coupling does not vary over the moduli space.
This implies that the corresponding IIB background should have constant
$\tau$. Indeed, in section \ref{IIB} we showed
that the 7-brane background for this configuration is very similar to the
background for the conformal theory. As in the conformal case, the 7-branes
do not bend and are coincident with the O7-planes. The RR charge of the 7-brane
is cancelled
locally by the O7-planes, so we expect that the type IIB string coupling is constant.
Similarly, the gravitational field of the 7-branes cancels against that of the
orientifold planes. Thus it appears that the closed string sector is not
affected by this deformation. The only difference between the conformal and the
non-conformal case is in the open string sector, namely in the gauge connection on the
7-branes. In the conformal case it is trivial, while in the non-conformal
case it is a flat connection which breaks the $SU(4)_7\times SU(4)_{7'}$ group to a
diagonally embedded $SU(2)\times SU(2)$. To summarize, the deformation of the 7-brane
background that leads to the non-conformal theory changes the properties of the
theory on the 7-branes, but it appears not to change the closed string sector.
To find a supergravity dual for this non-conformal theory, we need to repeat
the analysis above with the new 7-brane background. Since the closed string sector
is unchanged, the spectrum of the bulk modes should be the same as before.
The matter content of the conformal and the non-conformal theory differ only
in the number of flavors and their coupling to the bifundamentals. Therefore
both theories have the same spectrum of operators that do not transform under
the 7-brane groups. Thus it appears that the dimensions of all chiral primaries
uncharged with respect to the flavor group are the same as in the conformal case,
i.e., canonical. If antisymmetric tensors and bifundamentals have zero anomalous
dimensions then the vanishing of the beta-functions, \eq{beta2}, requires that
the fundamentals have dimension $1/2$. This is actually the lowest dimension
for the fundamental allowed by unitarity. To show that this assignment of dimensions,
or equivalently of R-charges, agrees with supergravity we would have to show that the KK
reduction of the 7-brane theory with the singular flat connection switched on, reproduces
the expected dimensions of the chiral primaries that involve the fundamentals.
Unfortunately we do not know how to analyze the excitations of the impurity theory
around nontrivial vacua, so we cannot check that our solution is consistent. Nevertheless,
we get a definite prediction for the infrared dimensions of all fields. It would be
interesting to confirm the answer by directly analyzing the perturbative expansion of the
non-conformal theory at large $N$.
\acknowledgements
It is a pleasure to thank O.~Aharony, E.~Gimon, J.~Maldacena, G.~Moore,
E.~Katz, and E.~Witten
for helpful discussions. M.G.~would like to thank the Institute for Advanced
Study for hospitality while this work was in progress.
The work of M.G~was supported in part by DOE grants \#DF-FC02-94ER40818 and
\#DE-FC-02-91ER40671, while
that of A.K. by DOE grant \#DE-FG02-90ER40542.
{\tighten
|
\section{Introduction}
The equivalence principle was postulated by Einstein as a
foundation stone for general relativity. The equivalence
principle stipulates that the long-range gravitational
interaction is entirely described by a universal coupling of
``matter'' (leptons, quarks, gauge fields and Higgs fields) to a
(dynamical) second-rank symmetric tensor field $g_{\mu \nu}
(x^{\lambda})$, replacing everywhere in the matter Lagrangian the
usual, kinematical, special relativistic (Minkowski) metric
$\eta_{\mu \nu}$. This principle assumes that all the
non-gravitational (dimensionless) coupling constants of matter
(gauge couplings, CKM mixing angles, mass ratios,$\ldots$) are
non-dynamical, i.e. take (at least at large distances) some fixed
(vacuum expectation) values, independently of where and when, in
spacetime, they are measured. Two of the best experimental tests
of the equivalence principle are:
(i) tests of the universality of free fall, i.e. the fact that
all bodies fall with the same acceleration in an external
gravitational field; and
(ii) tests of the ``constancy of the constants''.
Laboratory experiments (due notably, in our century, to
E\"otv\"os, Dicke, Braginsky and Adelberger) have verified the
universality of free fall to the $10^{-12}$ level. For instance,
the fractional difference in free fall acceleration of Beryllium
and Copper samples was found to be~\cite{Su94}
\begin{equation}
\left( \frac{\Delta a}{a} \right)_{\rm Be \, Cu} = (-1.9 \pm
2.5) \times 10^{-12} \, . \label{eq:adel}
\end{equation}
The Lunar Laser Ranging experiment~\cite{LLR} has also verified that the Moon
and the Earth fall with the same acceleration toward the Sun to better than one
part in $10^{12}$
\begin{equation}
\left( \frac{\Delta a}{a} \right)_{\rm Moon \, Earth} = (-3.2 \pm
4.6) \times 10^{-13} \, . \label{eq:llr}
\end{equation}
On the other hand, a recent reanalysis of the Oklo phenomenon (a
natural fission reactor which operated two billion years ago in
Gabon, Africa) gave a very tight limit on a possible time
variation of the fine-structure ``constant'', namely~\cite{DD96}
\begin{equation}
-0.9 \times 10^{-7} < \frac{e_{\rm Oklo}^2 - e_{\rm now}^2}{e^2}
< 1.2 \times 10^{-7} \, , \label{eq:oklo1}
\end{equation}
\begin{equation}
-6.7 \times 10^{-17} \, {\rm yr}^{-1} < \frac{d}{dt} \ {\rm
ln} \ e^2 < 5.0 \times 10^{-17} \, {\rm yr}^{-1} \, .
\label{eq:oklo2}
\end{equation}
The tightness of the experimental limits (\ref{eq:adel})--(\ref{eq:oklo2}) might
suggest
to apply Occam's razor and to declare that the equivalence
principle must be exactly enforced. However, the theoretical
framework of modern unification theories, and notably string
theory, suggest that the equivalence principle must be violated.
Even more, the type of violation of the equivalence principle
suggested by string theory is deeply woven into the basic fabric
of this theory. Indeed, string theory is a very ambitious attempt
at unifying all interactions within a consistent quantum
framework. A deep consequence of string theory is that
gravitational and gauge couplings are unified. In intuitive
terms, while Einstein proposed a framework where geometry and
gravitation were becoming united as a dynamical field $g_{\mu \nu}
(x)$, i.e. a soft structure influenced by the presence of matter,
string theory extends this idea by proposing a framework where
geometry, gravitation, gauge couplings, and gravitational
couplings all become soft structures described by interrelated
dynamical fields. A symbolic equation expressing this softened,
unified structure is
\begin{equation}
g_{\mu \nu} (x) \sim g^2 (x) \sim G(x) \, . \label{eq:ggg}
\end{equation}
It is conceptually pleasing to note that string theory proposes
to render dynamical the structures left rigid (or
kinematical) by general relativity. Technically,
Eq.~(\ref{eq:ggg}) refers to the fact that string theory (as well
as Kaluza-Klein theories) predicts the existence, at a
fundamental level, of scalar partners of Einstein's tensor field
$g_{\mu \nu}$, the model-independent ``dilaton'' field $\Phi
(x)$, and various ``moduli fields''. The dilaton field, notably,
plays a crucial role in string theory in that it determines the
basic ``string coupling constant'' $g_s = e^{\Phi (x)}$, which
determines in turn the (unified) gauge and gravitational coupling
constants $g \sim g_s$, $G \propto g_s^2$, as exemplified by the
low-energy effective action
\begin{equation}
L_{\rm eff} = e^{-2\Phi} \left[ \frac{R(g)}{\alpha'} +
\frac{4}{\alpha'} \, (\nabla \Phi)^2 - \frac{1}{4} \, F_{\mu
\nu}^2 - i \overline{\psi} \, D \psi - \ldots \right] \, .
\label{eq:eff}
\end{equation}
A softened structure of the type of Eq.~(\ref{eq:ggg}), embodied
in the effective action (\ref{eq:eff}), implies a deep violation
of Einstein's equivalence principle. Bodies of different nuclear
compositions fall with different accelerations because, for
instance, the part of the mass of nucleus $A$ linked to the
Coulomb interaction of the protons depends on the space-variable
fine-structure constant $e^2 (x)$ in a non-universal,
composition-dependent manner. This raises the problem of the
compatibility of the generic string prediction (\ref{eq:ggg})
with experimental tests of the equivalence principle, such as
Eqs.~(\ref{eq:adel}), (\ref{eq:llr}) or (\ref{eq:oklo2}). It is
often assumed that the softness (\ref{eq:ggg}) applies only at
short distances, because the dilaton and moduli fields are likely
to acquire a non zero mass after supersymmetry breaking. However,
a mechanism has been proposed~\cite{DP} to reconcile
in a natural manner the existence of a {\it massless} dilaton (or
moduli) field as a
fundamental partner of the graviton field $g_{\mu \nu}$ with
the current level of precision $(\sim 10^{-12})$ of
experimental tests of the equivalence principle. In the
mechanism of~\cite{DP} (see also~\cite{DN} for
metrically-coupled scalars) the very small couplings necessary
to ensure a near universality of free fall, $\Delta a/a <
10^{-12}$, are dynamically generated by the expansion of the
universe, and are compatible with couplings ``of order unity''
at a fundamental level.
The aim of the present paper is to emphasize the rich
phenomenological consequences of long-range dilaton-like fields, and the fact
that high-precision clock experiments might contribute to searching for, or
constraining, their existence.
More precisely, the basic question we wish to address here is the
following: given
the existing experimental tests of gravity, and given the
currently favored theoretical framework, can high-precision
clock experiments probe interesting theoretical possibilities
which remain yet unconstrained ? In addressing this question
we wish to assume, as theoretical framework, the class of
effective field theories suggested by string theory.
For historical completeness, let us mention that the theoretical
framework which has been most considered in the phenomenology of
gravitation, i.e. the class of
``metric'' theories of gravity~\cite{W81}, which includes most
notably the ``Brans-Dicke''-type tensor-scalar theories,
appears, from a modern perspective, as being rather artificial.
This is good news because
the phenomenology of ``non metric'' theories is richer and
offers new possibilities for clock experiments. Historically,
the restricted class of ``metric'' theories was introduced in
1956 by Fierz~\cite{F56} to prevent, in an {\it ad hoc} way,
too violent a conflict between experimental tests of the
equivalence principle and the existence of a scalar
contribution to gravity as suggested by the theories of
Kaluza-Klein~\cite{KK} and Jordan~\cite{J}. Indeed, Fierz was
the first one to notice that a Kaluza-Klein scalar would
generically strongly violate the equivalence principle. He
then proposed to restrict artificially the couplings of the
scalar field to matter so as to satisfy the equivalence
principle. The restricted class of
equivalence-principle-preserving couplings introduced by Fierz
is now called ``metric'' couplings. Under the aegis of Dicke,
Nordtvedt, Thorne and Will a lot of attention has been given
to ``metric'' theories of gravity\footnote{Note, however, that
Nordtvedt, Will, Haugan and others (for references see~\cite{W81}) studied
conceivable phenomenological
consequences of generic ``non metric'' couplings, without
using a motivated field-theory
framework to describe such couplings.}, and notably to their
quasi-stationary-weak-field phenomenology (``PPN framework'',
see, e.g.,~\cite{W81}).
\section{Generic effective theory of a long-range dilaton}
Motivated by string theory, we consider the generic class of
theories containing a long-range dilaton-like scalar field
$\varphi$. The effective Lagrangian describing these theories
has the form (after a conformal transformation to the ``Einstein
frame''):
\begin{eqnarray}
L_{\rm eff} &=& \frac{1}{4q} R(g_{\mu\nu}) - \frac{1}{2q} \
(\nabla \varphi)^2 - \frac{1}{4e^2 (\varphi)} \ (\nabla_{\mu}
A_{\nu} - \nabla_{\nu} A_{\mu})^2 \nonumber \\
&-& \sum_A \ \left[\overline{\psi}_A \,
\gamma^{\mu} (\nabla_{\mu} -iA_{\mu}) \psi_A + m_A (\varphi)
\, \overline{\psi}_A \psi_A \right] + \cdots \label{eq:01}
\end{eqnarray}
Here, $q\equiv 4\pi \, \overline G$ where $\overline G$ denotes
a bare Newton's constant, $A_{\mu}$ is the electromagnetic
field, and $\psi_A$ a Dirac field describing some fermionic
matter. At the low-energy, effective level (after the breaking
of $SU(2)$ and the confinement of colour), the coupling of the
dilaton $\varphi$ to matter is described by the
$\varphi$-dependence of the fine-structure ``constant'' $e^2
(\varphi)$ and of the various masses $m_A (\varphi)$. Here,
$A$ is a label to distinguish various particles. [A deeper
description would include more coupling functions, e.g.
describing the $\varphi$-dependences of the $U(1)_Y$,
$SU(2)_L$ and $SU(3)_c$ gauge coupling ``constants''.]
The strength of the coupling of the dilaton $\varphi$ to the
mass $m_A (\varphi)$ is given by the quantity
\begin{equation}
\alpha_A \equiv \frac{\partial \ {\rm ln} \ m_A
(\varphi_0)}{\partial \ \varphi_0} \, , \label{eq:02}
\end{equation}
where $\varphi_0$ denotes the ambient value of $\varphi (x)$
(vacuum expectation value of $\varphi (x)$ around the mass
$m_A$, as generated by external masses and cosmological
history). For instance, the usual PPN parameter $\gamma -1$
measuring the existence of a (scalar) deviation from the pure
tensor interaction of general relativity is given
by~\cite{DEF},~\cite{DP}
\begin{equation}
\gamma -1 = -2 \ \frac{\alpha_{\rm had}^2}{1+\alpha_{\rm had}^2}
\, , \label{eq:03}
\end{equation}
where $\alpha_{\rm had}$ is the (approximately universal)
coupling (\ref{eq:02}) when $A$ denotes any (mainly) hadronic
object.
The Lagrangian (\ref{eq:01}) also predicts (as discussed
in~\cite{DP}) a link between the coupling strength (\ref{eq:02})
and the violation of the universality of free fall:
\begin{equation}
\frac{a_A -a_B}{\frac{1}{2} (a_A + a_B)} \simeq (\alpha_A
-\alpha_B) \alpha_E \sim -5\times 10^{-5} \, \alpha_{\rm
had}^2 \, . \label{eq:04}
\end{equation}
Here, $A$ and $B$ denote two masses falling toward an external
mass $E$ (e.g. the Earth), and the numerical factor $-5 \times
10^{-5}$ corresponds to $A= {\rm Be}$ and $B= {\rm Cu}$. The
experimental limit Eq.~(\ref{eq:adel})
shows that the (mean hadronic) dilaton coupling strength is
already known to be very small:
\begin{equation}
\alpha_{\rm had}^2 \lsim 10^{-7} \, . \label{eq:06}
\end{equation}
Free fall experiments, such as Eq.~(\ref{eq:adel}) or the
comparable Lunar Laser Ranging constraint Eq.~(\ref{eq:llr}),
give the
tightest constraints on any long-range dilaton-like coupling.
Let us mention, for comparison, that solar-system measurements
of the PPN parameters (as well as binary pulsar measurements)
constrain the dilaton-hadron coupling to $\alpha_{\rm had}^2 <
10^{-3}$ (recently announced VLBI measurements improve this
constraint to the $2 \times 10^{-4}$ level), while the best
current constraint on the time
variation of the fine-structure ``constant'' (deduced from the
Oklo phenomenon), namely Eq.~(\ref{eq:oklo2}),
yields from Eq. (\ref{eq:16}) below, $\alpha_{\rm had}^2 \lsim
3 \times 10^{-4}$. [For an updated review of experimental tests
of gravity, see the chapter 14 of the Review of Particle Physics,
available on http://pdg.lbl.gov/]
To discuss the probing power of clock experiments, we need
also to introduce other coupling strengths, such as
\begin{equation}
\alpha_{\rm EM} \equiv \frac{\partial \ {\rm ln} \ e^2
(\varphi_0)}{\partial \ \varphi_0} \, , \label{eq:07}
\end{equation}
measuring the $\varphi$-variation of the electromagnetic (EM)
coupling constant\footnote{Note that we do not use the
traditional notation $\alpha$ for the fine-structure constant
$e^2 / 4\pi \hbar c$. We reserve the letter $\alpha$ for
denoting various dilaton-matter coupling strengths. Actually,
the latter coupling strengths are analogue to $e$ (rather than
to $e^2$), as witnessed by the fact that observable deviations
from Einsteinian predictions are proportional to products of
$\alpha$'s, such as $\alpha_A \alpha_E$, $\alpha_{\rm had}^2$,
etc$\ldots$}, and
\begin{equation}
\alpha_A^{A^*} \equiv \frac{\partial \ {\rm ln} \ E_A^{A^*}
(\varphi_0)}{\partial \ \varphi_0} \, , \label{eq:08}
\end{equation}
where $E_A^{A^*}$ is the energy difference between two atomic
energy levels.
In principle, the quantity $\alpha_A^{A^*}$ can be expressed
in terms of more fundamental quantities such as the ones
defined in Eqs. (\ref{eq:02}) and (\ref{eq:07}). For instance,
in an hyperfine transition
\begin{equation}
E_A^{A^*} \propto (m_e \, e^4) \ g_I \ \frac{m_e}{m_p} \ e^4 \
F_{\rm rel} (Z e^2) \, , \label{eq:09}
\end{equation}
so that
\begin{equation}
\alpha_A^{A^*} \simeq 2 \, \alpha_e -\alpha_p + \alpha_{\rm EM}
\left( 4+\frac{d \ {\rm ln} \ F_{\rm rel}}{d \ {\rm ln} \ e^2}
\right) \, . \label{eq:10}
\end{equation}
Here, the term $F_{\rm rel} (Z e^2)$ denotes the relativistic
(Casimir) correction factor~\cite{Casimir}. Moreover, in any
theory incorporating gauge unification one expects to have the
approximate link~\cite{DP}
\begin{equation}
\alpha_A \simeq \left( 40.75 - {\rm ln} \ \frac{m_A}{1 \ {\rm
GeV}} \right) \ \alpha_{\rm EM} \, , \label{eq:11}
\end{equation}
at least if $m_A$ is mainly hadronic.
\section{Clock experiments and dilaton couplings}
The coupling parameters introduced above allow one to describe
the deviations from general relativistic predictions in most
clock experiments~\cite{TD}. Let us only mention some simple
cases.
First, it is useful to distinguish between ``global'' clock
experiments where one compares spatially distant clocks, and
``local'' clock experiments where the clocks being compared
are next to each other. The simplest global clock experiment
is a static redshift experiment comparing (after transfer by
electromagnetic links) the frequencies of the same transition
$A^* \rightarrow A$ generated in two different locations ${\bf
r}_1$ and ${\bf r}_2$. The theory of Section 2 predicts a
redshift of the form (we use units in which $c=1$)
\begin{equation}
\frac{\nu_A^{A^*} ({\bf r}_1)}{\nu_A^{A^*} ({\bf r}_2)} \simeq
1 + (1 + \alpha_A^{A^*} \, \alpha_E) \ (\overline{U}_E ({\bf
r}_2) - \overline{U}_E ({\bf r}_1)) \, , \label{eq:12}
\end{equation}
where
\begin{equation}
\overline{U}_E \, ({\bf r}) = \frac{\overline G \, m_E}{r}
\label{eq:13}
\end{equation}
is the {\it bare} Newtonian potential generated by the
external mass $m_E$ (say, the Earth). Such a result has the
theoretical disadvantage of depending on other experiments for
its interpretation. Indeed, the {\it bare} potential
$\overline{U}_E$ is not directly measurable. The measurement
of the Earth potential by the motion of a certain mass $m_B$
gives access to $(1+\alpha_B \, \alpha_E) \ \overline{U}_E \,
({\bf r})$. The theoretical significance of a global clock
experiment such as (\ref{eq:12}) is therefore fairly indirect,
and involves other experiments and other dilaton couplings. One
can generalize (\ref{eq:12}) to a more general, non static
experiment in which different clocks in relative motion are
compared. Many different ``gravitational potentials'' will
enter the result, making the theoretical significance even
more involved.
A conceptually simpler (and, probably, technologically less
demanding) type of experiment is a differential, ``local''
clock experiment. Such ``null'' clock experiments have been
proposed by Will~\cite{W81} and first performed by Turneaure
et al.~\cite{T83}. The theoretical significance of such
experiments within the context of dilaton theories is much
simpler than that of global experiments. For instance if
(following the suggestion of~\cite{PTM}) one locally compares
two clocks based on hyperfine transitions in alkali atoms with
different atomic number $Z$, one expects to find a ratio of
frequencies
\begin{equation}
\frac{\nu_A^{A^*} ({\bf r})}{\nu_B^{B^*} ({\bf r})} \simeq
\frac{F_{\rm rel} (Z_A \, e^2 (\varphi_{\rm loc}))}{F_{\rm
rel} (Z_B \, e^2 (\varphi_{\rm loc}))} \, , \label{eq:14}
\end{equation}
where the local, ambient value of the dilaton field
$\varphi_{\rm loc}$ might vary because of the (relative)
motion of external masses with respect to the clocks
(including the effect of the cosmological expansion). The
directly observable fractional variation of the ratio
(\ref{eq:14}) will consist of two factors:
\begin{equation}
\delta \ {\rm ln} \ \frac{\nu_A^{A^*}}{\nu_B^{B^*}} = \left[
\frac{\partial \ {\rm ln} \ F_{\rm rel} (Z_A \, e^2)}{\partial
\ {\rm ln} \ e^2} - \frac{\partial \ {\rm ln} \ F_{\rm rel}
(Z_B \, e^2)}{\partial \ {\rm ln} \ e^2} \right] \times \delta
\ {\rm ln} \ e^2 \, . \label{eq:15}
\end{equation}
The ``sensitivity'' factor in brackets due to the
$Z$-dependence of the Casimir term can be made of order
unity~\cite{PTM}, while the fractional variation of the
fine-structure constant is expected in dilaton theories to be
of order~\cite{DP},~\cite{TD}
\begin{eqnarray}
\delta \ {\rm ln} \ e^2 (t) &=& -2.5 \times 10^{-2} \
\alpha_{\rm had}^2 \ U(t) \nonumber \\
&-& 4.7 \times 10^{-3} \ \kappa^{-1/2} ({\rm tan} \ \theta_0)
\ \alpha_{\rm had}^2 \ H_0 (t-t_0) \, . \label{eq:16}
\end{eqnarray}
Here, $U(t)$ is the value of the externally generated
gravitational potential at the location of the clocks, and
$H_0 \simeq 0.5 \times 10^{-10} \ {\rm yr}^{-1}$ is the Hubble
rate of expansion. [The factor $\kappa^{-1/2} \ {\rm tan} \
\theta_0$ is expected to be $\sim 1$.]
The (rough) theoretical prediction (\ref{eq:16}) allows one to
compare quantitatively the probing power of clock experiments
to that of equivalence principle tests. Let us
(optimistically) assume that clock stabilities of order
$\delta \nu / \nu \sim 10^{-17}$ (for the relevant time scale)
can be achieved. A differential {\it ground} experiment (using
the variation of the Sun's potential due to the Earth
eccentricity) would probe the level $\alpha_{\rm had}^2 \sim
3\times 10^{-6}$. A geocentric satellite differential
experiment could probe $\alpha_{\rm had}^2 \sim 5\times
10^{-7}$. These levels are impressive (compared to present
solar-system tests of the PPN parameter $\gamma$ giving the
constraint $\alpha_{\rm had}^2 \simeq (1-\gamma) / 2 <
2 \times 10^{-4}$), but are not as good as the present
equivalence-principle limit (\ref{eq:06}). To beat the level
(\ref{eq:06}) one needs to envisage an heliocentric
differential clock experiment (a few solar radii probe within
which two hyper-stable clocks are compared). Such an
experiment could, according to Eq. (\ref{eq:16}), reach the
level $\alpha_{\rm had}^2 \sim 10^{-9}$. [Let us also note
that a gravitational time delay global experiment using clocks
beyond the Sun as proposed by C. Veillet (SORT concept) might
(optimistically) probe the level $\alpha_{\rm had}^2 \sim
10^{-7}$.] It is, however, to be noted that a much refined test
of the equivalence principle such as STEP (Satellite Test of
the Equivalence Principle) aims at measuring $\Delta a/a \sim
10^{-18}$ which corresponds to the level $\alpha_{\rm had}^2
\sim 10^{-14}$, i.e. five orders of magnitude better than any
conceivable clock experiment.
\section{Conclusions}
In summary, the main points of the present contribution are:
\begin{enumerate}
\item[$\bullet$] Independently of any theory, the result
(\ref{eq:oklo2}) of a recent reanalysis of the Oklo
phenomenon~\cite{DD96} gives a motivation, and a target, for
improving
laboratory clock tests of the time variation of the
fine-structure constant $e^2$ (which are at the $3.7 \times
10^{-14} \ {\rm yr}^{-1}$ level~\cite{PTM}).
\item[$\bullet$] Modern unification theories, and especially
string theory, suggest the existence of new
gravitational-strength fields, notably scalar ones
(``dilaton'' or ``moduli''), whose couplings to matter violate
the equivalence principle. These fields would induce a
spacetime variability of the coupling constants of physics
(such as the fine-structure constant). High-precision clock
experiments are excellent probes of such a possibility.
\item[$\bullet$] The generic class of dilaton theories defined
in Section 2 provides a well-defined theoretical framework in
which one can discuss the phenomenological consequences of the
existence of a dilaton-like field. Such a theoretical
framework (together with some assumptions, e.g. about gauge
unification and the origin of mass hierarchy) allows one to
compare and contrast the probing power of clock experiments to
that of other experiments.
\item[$\bullet$] Local, differential clock experiments (of the
``null'' type of~\cite{T83}) appear as conceptually cleaner,
and technologically less demanding, probes of
dilaton-motivated violations of the equivalence principle than
global, absolute clock experiments (of the Gravity Probe A
type).
\item[$\bullet$] If we use the theoretical assumptions of
Section 2 to compare clock experiments to free-fall
experiments, one finds that one needs to send and intercompare
two ultra-high-stability clocks in near-solar orbit in order
to probe dilaton-like theories more deeply than {\it present}
free-fall experiments. Currently proposed improved satellite
tests of the equivalence principle would, however, beat any
clock experiment in probing even more deeply such theories.
\item[$\bullet$] At the qualitative level, it is, however,
important to note that clock experiments (especially of the
``global'', GPA type) probe different combinations of basic
coupling parameters than free-fall experiments. This is
visible in Eq. (\ref{eq:10}) which shows that $\alpha_A^{A^*}$
contains the leptonic quantity $\alpha_e = \partial \ {\rm ln}
\ m_{\rm electron} / \partial \ \varphi_0$ without any small
factor\footnote{Free-fall experiments couple predominantly to
hadronic quantities such as $\alpha_p =
\partial \ {\rm ln} \ m_{\rm proton} / \partial \ \varphi_0$,
and to Coulomb-energy effects proportional to $\alpha_{\rm
EM}$. The effect of the leptonic quantity $\alpha_e$ is down
by a small factor $\sim m_e / m_p \sim 1/1836$.}.
\end{enumerate}
\section*{References}
|
\section{Introduction}
Several years after the first demonstration \cite{R1}, optical homodyne
tomography has become a well established tool for measuring quantum
statistical properties of optical radiation. What is particularly
fascinating, this technique provides practical means to visualise the
measured quantum state in the form of the Wigner function. This success
is a result of combining a complete quantum mechanical measurement of
field quadratures with a filtered back-projection algorithm used in
medical imaging.
The purpose of this contribution is to trace some other analogies
between quantum state reconstruction and classical image processing,
with the motivation to develop novel numerical methods for quantum
tomography. Our interest will be focused on imperfect detection
\cite{R2,R3}, which has deleterious effects on quantum interference phenomena
exhibited by non-classical states \cite{R4}. As we will discuss in Sec.~2,
such effects can be related in the phase space representation to image
blurring. Restoration of blurred images is a well known problem in
classical imaging, and a number of methods has been developed for this
purpose. Specifically, we shall briefly describe in Sec.~3 the Richardson
algorithm \cite{R5} (known also in statistics as the expectation-maximization
algorithm \cite{R6}), which provides an effective iterative procedure to
perform image deblurring.
An interesting question is, whether classical deblurring methods can
be transferred to quantum tomography. We discuss this, in the case of
the Richardson algorithm, in Sec.~4. The answer is not straightforward:
the Richardson algorithm assumes positive definiteness of the original
undegraded image, and this condition is not satisfied by the quantum
mechanical Wigner function. We will show that this difficulty can be
overcome by expressing the Wigner function in terms of the phase space
displaced photon distribution. This yields an iterative algorithm for
reconstructing the Wigner function, which incorporates compensation for
detection losses in a numerically stable way \cite{R7}.
\section{Imperfect detection and image blurring}
Homodyne detection is a realization of the quantum mechanical measurement
of field quadratures only in the idealized loss-free limit.
In practice, a fraction of the signal field is always lost
due to the mode-mismatch and the non-unit efficiency of photodiodes.
The homodyne statistics collected using
a realistic setup is described by the distribution \cite{R8}:
\begin{equation}
\label{Eq:hxtheta}
h(x;\theta) =
\int_{-\infty}^{\infty} \mbox{d}x' \;
\frac{1}{\sqrt{\pi(1-\eta)}} \exp \left(
- \frac{(x-\sqrt{\eta}x')^{2}}{1-\eta}
\right)
\langle x'_{\theta} | \hat{\rho} | x'_{\theta} \rangle,
\end{equation}
where $\theta$ is the local oscillator phase, $\eta$ characterizes
the overall detector efficiency, and $\langle x'_{\theta} | \hat{\rho}
| x'_{\theta} \rangle$ denote diagonal elements of the density matrix
$\hat{\rho}$ in the quadrature basis.
Application of the back-projection transformation to $h(x;\theta)$
yields, instead of the Wigner function, a generalized phase space
quasidistribution function $P(q,p; s)$ with the ordering parameter
$s=-(1-\eta)/\eta$. This function can be expressed as a
convolution of the Wigner function $W(q,p)$ with a gaussian function:
\begin{equation}
\label{Eq:Palphas}
P(q,p; -|s|) = \int \mbox{d}q'\,\mbox{d}p' \; \frac{1}{\pi |s|}
\exp\left( - \frac{1}{|s|}[(q-q')^2+(p-p')^2 ]\right) W(q',p').
\end{equation}
Thus what we reconstruct from imperfect homodyne data,
is a blurred version of the Wigner function. The question is, whether we
can get rid of this blurring in numerical processing of experimental data.
A similar problem appears in the following classical context:
suppose we observe an image using an imperfect apparatus (for example
ill-matched glasses), which generates some blurring. Such blurring can be
described by a so-called point spread function specifying the shape
generated by a single point of the original image. The observed
degraded image is consequently
given by a convolution of the original image with
the point spread function. Using this language, we can assign the
following names to the terms of Eq.~(\ref{Eq:Palphas}):
\begin{center}
\begin{tabular}{rcl}
$W(q',p')$ & --- & original image \\
$ \displaystyle
\frac{1}{\pi |s|} \exp\left( - \frac{1}{|s|}
[(q-q')^2+(p-p')^2]\right)$
& --- & point spread function \\
$P(q,p; -|s|)$ & --- & degraded image
\end{tabular}
\end{center}
The common problem now is the restoration of the original image
from the degraded one, assuming that the point spread function
is known.
\section{Image restoration}
An analytical way to deconvolve Eq.~(\ref{Eq:Palphas}) is to apply
the Fourier transform, which maps a convolution onto a direct product.
Dividing both the sides by the Fourier-transformed point spread function
and evaluating the inverse Fourier transform thus yields an explicit
expression for the original image. However, this procedure is very
sensitive to statistical fluctuations and numerical truncation errors,
which makes its practical application a very delicate matter. These
problems have been noted also in the context of quantum tomography \cite{R9}.
The numerical instability of the Fourier deconvolution has led to the
development of techniques dedicated for image restoration. The
basic observation is that statistical noise does not allow us to specify
precisely the original image that was 'behind' the blurred data. In
principle, the measured degraded image could could be generated by a
variety of original images. However, comparing various original images we
intuitively expect that some of them were {\em more likely} to generate
the measured data than other ones. The maximum-likelihood methodology
quantifies this intuition, and selects as an estimate the original image
which maximizes the likelihood.
In order to discuss this idea in detail, we shall
consider a discretized version of Eq.~(\ref{Eq:Palphas}):
\begin{equation}
\label{Eq:LININPOS}
p_{\nu} = \sum_{n} A_{\nu n} w_{n},
\end{equation}
where $w_n$ is the original image,
$A_{\nu n}$ is the point spread function, and
$p_{\nu}$ is the degraded image. Note that this formulation
is more general compared to Eq.~(\ref{Eq:Palphas}), because
it allows the point spread function to be of different form
for each 'element' of the original image indexed with $n$.
The likelihood can be quantified using the function
\begin{equation}
{\cal L} = \sum_{\nu} p_{\nu} \ln \left(
\sum_{n} A_{\nu n} w_{n} \right)
\end{equation}
which has a rigorous derivation when the degraded image is observed as a
histogram of events governed by Poissonian statistics. The likelihood
function for quantum measurement has been discussed in Ref.~\cite{R10},
where its maximization has been proposed as a method for quantum state
estimation.
In classical imaging, it is natural to assume that $w_{n}$, as well as
$A_{\nu n}$ as a function of $\nu$ for each $n$, are positive definite
distributions with sum equal to one.
Under these assumptions, it is
possible to find the maximum of the likelihood function ${\cal L}$
via simple iteration:
\begin{equation}
w_{n}^{(i+1)} = \sum_{\nu} p_{\nu}
\frac{A_{\nu n} w_{n}^{(i)}}{\sum_{m} A_{\nu m} w_{m}^{(i)}},
\end{equation}
which is the essence of the Richardson algorithm for image
restoration~\cite{R5}. A simple heuristic derivation of this algorithm
can be found in Ref.~\cite{R7}.
Of course, the maximum-likelihood approach is not a magic wand solving
unconditionally the problem of image restoration. With increasing
blurring, the quality of the reconstructed image worsens, and the
convergence of the iterative algorithm may be very slow. In many cases,
however, it offers superior performance compared to the Fourier
deconvolution technique.
\section{Quantum tomography}
An obvious difficulty with applying the iterative restoration algorithm to
quantum tomography is that the object to be reconstructed in the quantum
case, i.e.\ the Wigner function, is not positive definite. Nevertheless,
there are some other quantum mechanical reconstruction problems, where
positivity constraints appear in a natural way. An interesting and
nontrivial example is determination of the photon number distribution
from random phase homodyne statistics \cite{R11}. The relation between the
phase-averaged homodyne statistics and the photon number distribution
$\varrho_n$ is given by
\begin{equation}
\frac{1}{2\pi}\int_{-\pi}^{\pi} \mbox{d}\theta \;
h(x;\theta) = \sum_{n} A_n(x) \varrho_n,
\end{equation}
where $A_{n}(x)$ describe contributions to the homodyne statistics
generated by different Fock states $|n\rangle$.
This formula, after discretization of $x$, is exactly of the form
assumed in Eq.~(\ref{Eq:LININPOS}).
Thus we arrive at the following formal analogy:
\begin{center}
\begin{tabular}{rcl}
$\varrho_n$ & --- & original image \\
$A_{n}(x)$ & --- & point spread function \\
$\displaystyle \frac{1}{2\pi}
\int_{-\pi}^{\pi}
\mbox{d}\theta \; h(x; \theta)$ & --- &
degraded image
\end{tabular}
\end{center}
which allows us to apply directly the iterative reconstruction algorithm
\cite{R12}. In this procedure, there is one {\em a priori} parameter: it is the
cut-off of the distribution $\varrho_n$ specifying the maximum number
of photons.
Reconstruction of the photon distribution may seem to be quite distant
from the starting point of our considerations, which was deblurring
of the Wigner function. However, let us recall that the Wigner function
can be represented as an alternating sum of the photon distribution
$\varrho_{n}(q,p)$ corresponding to the phase space displaced state
$\hat{D}^{\dagger}(q,p)\hat{\varrho}\hat{D}(q,p)$:
\begin{equation}
\label{Eq:Wqp}
W(q,p) = \frac{1}{\pi} \sum_{n=0}^{\infty} (-1)^{n}
\varrho_{n}(q,p)
\end{equation}
Obviously, we can apply this formula to evaluate the Wigner function
at $q=p=0$. What would be of interest, is the generalization of the
maximum-likelihood algorithm to determination of an arbitrarily displaced
photon distribution $\varrho_n(q,p)$. This would yield a numerically
stable procedure for reconstructing the Wigner function from homodyne
statistics, even in the case of the non-unit detection efficiency.
Surprisingly, this generalization is quite straightforward. The basic
observation is that the displacement transformation has a simple effect
on the homodyne statistics, shifting it by
$\sqrt{\eta}(q\cos\theta+p\sin\theta)$ for a given local oscillator
phase $\theta$. Consequently, we have the relation
\begin{equation}
\label{Eq:tomolininpos}
\frac{1}{2\pi}\int_{-\infty}^{\infty} \mbox{d}\theta \;
h(x+\sqrt{\eta} q \cos\theta + \sqrt{\eta}p \sin\theta ; \theta)
= \sum_{n} A_{n}(x) \varrho_{n}(q,p),
\end{equation}
which can be readily implemented in the iterative algorithm.
Thus, we have arrived at the following two-step algorithm for quantum
tomography: for a given phase space point $(q,p)$, construct the
phase-averaged histogram according to the right-hand side
of Eq.~(\ref{Eq:tomolininpos}), and iteratively reconstruct
$\varrho_n(q,p)$. Then, calculate the value of the Wigner function
according to Eq.~(\ref{Eq:Wqp}). In Fig.~1 we present Monte Carlo
simulated reconstruction of
the Wigner function for a Schr\"{o}dinger cat state detected using
a homodyne setup with the efficiency $\eta=90\%$.
\begin{figure}
\centerline{\setlength{\unitlength}{1pt}
\begin{picture}(325,250)
\put(0,0){\epsfig{file=kot.eps}}
\put(20,225){\small $W(q,p)$}
\put(120,0){\small $q$}
\put(235,25){\small $p$}
\end{picture}}
\bigskip
\caption{Fig.~1. Reconstruction of the Schr\"{o}dinger cat state
$|\Psi\rangle\propto|2i\rangle -|-2i\rangle$ from Monte Carlo simulated
homodyne experiment. The homodyne data consisted of $10^5$ events
generated for each of $64$ phases spaced uniformly between $0$ and
$\pi$. At each point of the grid, the displaced photon statistics
was obtained from $10^4$ iterations, starting from a flat distribution
for $0\le n \le 39$. In the simulations, the homodyne variable $x$ has
been discretized into 16000 bins over the range $-8\le x \le 8$.}
\end{figure}
The standard filtered back-projection algorithm used in quantum
tomography is based on the inverse Radon transform, whose integral
kernel is singular. Therefore, a regularization scheme is necessary
in processing experimental data. This aspect has a counterpart in the
maximum-likelihood algorithm. In this approach, we have the cut-off
for the photon distribution which can be regarded as a regularization
parameter. Its proper choice is an important matter. Setting it too
small perturbes the reconstructed photon distribution. On the other hand,
the larger number of $\varrho_n$s, the slower iterations converge. The
expected shape of the photon distribution can be quite easily predicted,
if we roughly know the region of the phase space where the Wigner function
is localized. For this purpose it is useful to recall the semiclassical
picture of projections on Fock states as rings in the phase space
characterized by the radius $\sqrt{2n}$. The photon distribution is
nonzero over the range of $n$ for which the corresponding rings overlap
with the localization region for the Wigner function.
Truncation of the photon distribution can be introduced as a
regularization scheme also in the standard linear reconstruction
approach. In such a scheme, the Wigner function would be evaluated
from a finite part of the photon distribution reconstructed using the
pattern function technique. However, properties of the reconstructed
photon distribution make this method very sensitive to statistical
noise. This can be straightforwardly seen in the most regular case of
$\eta=1$. For large $n$, the error of $\varrho_n$ tends to a fixed nonzero
value \cite{R13}, and moreover deviations of consecutive $\varrho_n$s are
strongly anticorrelated. Consequently, the alternating sum defined in
Eq.~(\ref{Eq:Wqp}) accumulates the statistical uncertainty of the photon
distribution \cite{R14}.
Let us also note that in principle we could apply the restoration
algorithm to homodyne histograms described by Eq.~(\ref{Eq:hxtheta}),
in order to obtain deblurred quadrature distributions $\langle x_\theta
| \hat{\varrho} | x_\theta \rangle$. In this case statistical
fluctuations would play a much more significant role. The advantage of
using Eq.~(\ref{Eq:tomolininpos}) is that we use the whole available
sample of experimental data to determine a relatively small number of
parameters $\varrho_{n}(q,p)$.
\medskip
\noindent{\bf Acknowledgements}
I would like to thank Prof.\ G.~Mauro D'Ariano for his hospitality
during my stay in Pavia supported by INFM, and Prof.\ Krzysztof
W\'{o}dkiewicz for comments on the manuscript. I have benefited a lot
from discussions with Zdenek Hradil on maximum-likelihood methods in
quantum state reconstruction. This research is supported by
Komitet Bada\'{n} Naukowych, grant 2P03B~013~15.
|
\section{Introduction.}
One of the most popular approaches used to determine the small amplitude
envelope
soliton solutions in non linear models is the well-known
Multiple Scale
Expansion (MSE) technique \cite{rem}.
This technique amounts to expanding the equations of motion on
different time and space scales
looking for wave-packet-like solutions; a wave packet is
a superposition of plane waves whose frequencies and wave vectors lie
in a narrow band, and it can be conveniently described by a plane wave
with an amplitude that varies slowly in space and time.
Increasing progressively the time and space
scales one determines in a first step
the carrier wave as a phonon mode of the linearized system,
then deduces the partial differential equation that identifies the envelope
phase velocity with the wave packet group velocity, and finally derives
the NLS equation for the envelope whose diffusion coefficient is in fact
the wave packet group velocity dispersion.
An alternative method, commonly used in optics,
consists in expanding the dispersion relations with
respect to the carrier frequency and then in building at each order of this expansion
an operator that acts on the envelope function
\cite{opt}.
MSE has been successfully applied to various non linear systems
with scalar fields and the corresponding method adopted in optics
has permitted the study of optical solitons in fiber.
In this latter case one deals with eletric field components which are
not coupled
at the linear order of the Maxwell equations.
However, many non linear models of interest
involve vectorial fields with coupled components at the linear
order that give rise to dispersion relations with more than one branch:
classical examples are given by
multi-atomic lattices, or by lattices in which the mass at
each site can move in a multidimensional space.
In this work, we show how to find small amplitude envelope soliton solutions
in such vectorial lattices problems.
The main difficulty with respect
to the scalar case is to determine the relative amplitudes of the
different components of the field.
We will perform a perturbative expansion, around one
wavenumber, of the linear system that gives the dispersion relations
and the linear eigenmodes; then
we will introduce an operator
formalism analogous to that used in optics to obtain, from this
expansion and from the nonlinear terms,
the MSE
equations, up to the NLS one.
An application of the method presented in this work
can be found in a forthcoming paper \cite{noi}, where
the envelope
soliton solutions of an helicoidal DNA model
described by a radial and an angular degree of freedom for each site
are derived.
\section {Wave-packet in linear vectorial lattices}
The NLS equation is obtained when a weak dispersion is balanced by a
weak nonlinearity. In order to characterize the dispersion, let us
first restrict our attention to the linear part of the system of interest.
We start with a one dimensional vectorial linear lattice model
given by the equations of motion:
\begin{equation}
\label{emoto}
\frac{\partial^2 E(n,\alpha,t)}{\partial t^2 }=
- \sum_{n',\alpha'} J (n-n',\alpha,\alpha') E(n',\alpha',t)
\end{equation}
where $n,n'$ are the site indices, $\alpha,\alpha'$ are the
indices that label the components of the vectorial field
$ E(n,\alpha,t)$, and
$J (n-n',\alpha,\alpha')$ are the force constants depending on
$ n-n'$ for translationally invariant systems.
Looking for plane wave solutions of the form
\begin{equation}
A \vec{V}_l(q) e^{i(q n-\omega_l(q)t)} + c.c.
\end{equation}
where $A$ is the wave amplitude, the equation of motion is mapped to the
operator equation in the wave numbers space:
\begin{equation}
\label{d}
(\hat{J}(q)- \omega^{2}_{l}(q))\vec{V}_{l}(q) =0
\end{equation}
where $\hat{J}(q)$ is the Fourier transform of the matrix $\hat{J}(n-n')$.
The index $l$ runs from 1 to the number of components of the vectorial
field $E(n,\alpha,t)$; the eigenvalues functions $\omega^{2}_{l}(q)$
give the branches of the dispersion relation;
the normal modes $\vec{V}_l(q)$ are the orthonormalized eigenvectors of
the matrix $\hat{J}(q)- \omega^{2}_{l}(q) $.
In order to investigate the dispersion, we now consider a wave-packet-like
solution, {\em i.e.} a superposition of plane waves with wave numbers
in a small interval:
\begin{equation}
\label{wpuno}
\vec{E}_l(n,t)=\int_{q_0 - \Delta q}^{q_0 + \Delta q}
A(q) \, \vec{V}_{l}(q) \,
e^{i(q n-\omega_l(q)t)} \; dq +c.c.
\end{equation}
For each $q$ contributing to the wave packet the system of equations
(\ref{d}) must
be fulfilled. The weakly dispersive case
is obtained by considering only small deviations of $q$ with
respect to the wavector $q_0$ corresponding to the center of the
wavepacket. To measure this deviation, the wavevector $q$ is written $
q = q_0 + \epsilon q_1$, where $\epsilon \ll 1$.
Eq.~(\ref{d}) is
solved, $\forall q_1$ in the integration
range, by a perturbative technique.
The operator $\hat{J}(q_0+\epsilon q_1)$ can be expanded in Taylor series as
$\hat{J}(q_0)+\epsilon \hat{J}'(q_0)q_1+\epsilon^2
\hat{J}''(q_0)q_{1}^{2}/2 + \ldots$. The quantities
$\epsilon \hat{J'}(q_0)q_1$, $\epsilon^2 \hat{J''}(q_0)q_1^2/2$
are small perturbations
with respect to unperturbed operator $\hat{J}(q_0)$,
whose eigenvalues are $\omega^2_l=\omega^2_l(q_0)$ and whose
eigenvectors $\vec{V}_{l}=\vec{V}_{l}(q_0)$ constitute a complete basis.
According to standard perturbation theory \cite{coh} we
write the expansions of the eigenvectors and eigenvalues,
\begin{eqnarray}
\label{v}
\vec{V}_{l}(q_0+\epsilon q_1) &=&\vec{V}_{l}+ \epsilon \vec{V}^{(1)}_{l}q_1
+ \epsilon^2 \vec{V}^{(2)}_{l}q_1^2/2 +\ldots \\
\label{omega}
\omega_l(q_0+\epsilon q_1) &=&\omega_l+\epsilon
\omega^{(1)}_l q_1 +
\epsilon^2 \omega^{(2)}_l q_1^2/2+ \ldots
\end{eqnarray}
Equation (\ref{d}) has to be solved at each expansion order:
\begin{eqnarray}
\label{o1} \text{at order $\epsilon^0$ : }\quad
\hat{J}\vec{V}_l &=& \omega^2_l \vec{V}_l \\
\label{o2}
\text{at order $\epsilon^1$ : }\quad
(\hat{J} \vec{V}^{(1)}_l+\hat{J}' \vec{V}_l)\;q_1 &=&
(2\omega_l \omega^{(1)}_l \vec{V}_l+ \omega^2_l \vec{V}^{(1)}_l)\;q_1 \\
\label{o3}
\text{at order $\epsilon^2$ : }\quad
(\hat{J}' \vec{V}^{(1)}_l + \frac{1}{2} \hat{J}'' \vec{V}_l+
\frac{1}{2} \hat{J} \vec{V}^{(2)}_l)\;q_1^2 &=&
{(\omega^{(1)}}^2_l \vec{V}_l +\omega_l \omega^{(2)}_l\vec{V}_l
+2\omega_l \omega^{(1)}_l \vec{V}^{(1)}_l+
\frac{\omega^2_l}{2} \vec{V}^{(2)}_l)\;q_1^2 \,.
\end{eqnarray}
At order $\epsilon^0$,
one solves the unperturbed problem determining $\vec{V}_l$
and $\omega_l$. At order $\epsilon$, one determines $\vec{V}^{(1)}_l$,
$\omega^{(1)}_l$: imposing to $\vec{V}^{(1)}_l$ to be orthogonal to $\vec{V}_l$ to
guarantee the normalization of $\vec{V}_l(q_0+\epsilon q_1)$ the scalar
product of (\ref{o2}) with $\vec{V}^*_m \;\; (m\neq l)$ gives
\begin{eqnarray}
\label{vp}
\vec{V}^{(1)}_l= \sum_{m\neq l} \alpha_m \vec{V}_m \\
\label{am}
\alpha_m= \frac{ {\vec{V}_m}^* \hat{J}' \vec{V}_l}{ \omega^2_l-\omega^2_m}
\;\;\;\;\;\;\; m \neq l
\end{eqnarray}
and that with $\vec{V}^*_l$ gives
\begin{equation}
\label{op}
\omega^{(1)}_l= \frac{ {\vec{V}_l}^* \hat{J}' \vec{V}_l}{2 \omega_l}\,.
\end{equation}
At order $\epsilon^2$, one determines $\omega^{(2)}_l$ by
multiplying (\ref{o3}) by $\vec{V}^*_l$
\begin{equation}
\label{os}
\omega^{(2)}_l= \frac{1}{\omega_l} \left({\vec{V}_l}^* \frac{\hat{J}''}{2} \vec{V}_l
-{\omega^{(1)}}^2_l+
\sum_{m\neq l} \frac{| {\vec{V}_m}^* \hat{J}' \vec{V}_l|^2} {
\omega^2_l-\omega^2_m} \right)\,.
\end{equation}
We assume, for sake of simplicity, that the eigenmodes of J
(see (\ref{d}))
are not degenerate, but the generalization of the degenerate
case is straightforward with the standard perturbation theory.
The phase of each component of (\ref{wpuno}) can be expanded around
the central wave number $q_0$, up to second order in $\epsilon q_1=q-q_0$
using the values of $\omega^{(1)}_l$ and $\omega^{(2)}_l$ determined above
\begin{equation}
\label{wp}
\vec{E}_l(n,t)= e^{i(q_0 n-\omega_l(q_0)t)}
\epsilon \int_{-\Delta q/\epsilon}^{+\Delta q/\epsilon}
A(q_0+ \epsilon q_1) \vec{V}_{l}(q_0+ \epsilon q_1)
\;
\exp{\left\{i \epsilon q_1(n-\omega^{(1)}_l(q_0)t)
-i \epsilon^2 \frac{q_1^2}{2}\omega^{(2)}_l(q_0)t\right\}}
\; dq_1+ c.c.
\end{equation}
Under this form, $\vec{E}_l(n,t)$ appears as a plane wave, henceforth
called the carrier wave, with an
amplitude that depends on space and time and which corresponds to
the integral of equation (\ref{wp}),
$\vec{E}_l(n,t) = \vec{F}(n,t) \; \exp[i(q_0 n-\omega_l(q_0)t)] +c.c.$.
The fact that $\omega^{(1)}_l(q_0)$,
$\omega^{(2)}_l(q_0)$ obey relations (\ref{op}) and (\ref{os}) ensures
that this wave packet is a solution of the original equation
(\ref{emoto}), up to the order of the various expansions.
In order to
extend the study to the nonlinear case it is useful to express these
conditions under the form of an equation in the space time coordinates
for the amplitude. Let us introduce the quantity
\begin{equation}
\label{ap}
A(n,t)= \int_{-\Delta q/\epsilon}^{+\Delta q/\epsilon} A(q_0+ \epsilon q_1)
\;
\exp{\left\{i \epsilon q_1(n-\omega^{(1)}_l(q_0)t)
-i \epsilon^2 \frac{q_1^2}{2}\omega^{(2)}_l(q_0)t\right\}}
\; dq_1
\end{equation}
Equation (\ref{ap}) shows that $A(n,t)$ slowly varies in space and
time. In the spirit of the multiple scale expansion, it is natural to
introduce the slow variables $x_1 = \epsilon n$, $t_1 = \epsilon t$
and $t_2 = \epsilon^2 t$ so that $A(n,t)$ can be written as
\begin{equation}
\label{ap1}
A(n,t) = A(x_1,t_1,t_2) =
\int_{-\Delta q/\epsilon}^{+\Delta q/\epsilon} A(q_0+ \epsilon q_1)
\;
\exp{\left\{i q_1( x_1-\omega^{(1)}_l(q_0)t_1)
-i \frac{q_1^2}{2}\omega^{(2)}_l(q_0)t_2\right\}}
\; dq_1
\end{equation}
or
\begin{equation}
\label{ap2}
A(x_1,t_1,t_2) = A(s_1,t_2) =
\int_{-\Delta q/\epsilon}^{+\Delta q/\epsilon} A(q_0+ \epsilon q_1)
\;
\exp{\left\{i q_1 s_1
-i \frac{q_1^2}{2}\omega^{(2)}_l(q_0)t_2\right\}}
\; dq_1
\end{equation}
with the introduction of the variable $s_1 = x_1 - \omega^{(1)}_l(q_0)t_1$
to switch to the frame moving at the group velocity of
the carrier wave.
Using the relation $(\partial A / \partial x_1) = (\partial A /
\partial s_1) = i<q_1> \equiv
\int_{-\Delta q/\epsilon}^{+\Delta q/\epsilon}
i q_1\; A(q_0+\epsilon q_1)
\;
\exp{\left\{i q_1 s_1
-i \frac{q_1^2}{2}\omega^{(2)}_l(q_0)t_2\right\}}
\; dq_1$ that derives directly from
Eqs.~(\ref{ap1}) and (\ref{ap2}), and the expansion (\ref{v}) of $V(q_0
+ \epsilon q_1)$, the amplitude $\vec{F}$ of the wave can be
expressed as a function of $A(s_1,t_2)$ by the relation
\begin{equation}
\label{f}
\vec{F}(x_1,t_1,t_2)= \epsilon
(\vec{V}_{l}-i
\epsilon \vec {V'_l} \frac{\partial }{\partial x_1}) A(x_1,t_1,t_2)
\; .
\end{equation}
{From} (\ref{ap1}) and (\ref{ap2}), we directly derive
the equations of motions of $A$ as a function of the slow space--time
variables:
\begin{equation}
\label{wp1}
\left( \frac{\partial A}{\partial t_1}+
\omega^{(1)}_l\frac{\partial A}{\partial x_1} \right) =0
\end{equation}
and
\begin{equation}
\label{ls}
i \frac{\partial A }{\partial t_2}+
\frac{\omega^{(2)}_l}{2}\frac{\partial^2 A}{\partial s_1^2}=0
\end{equation}
where $\omega^{(1)}_l$ and $\omega^{(2)}_l$ are then the group velocity and
the group velocity dispersion of the wave packet and determine
the peak velocity and the spread out of the envelope function.
Equation (\ref{f}) shows that
$\vec{V'}_{l}$ determines the first order correction to the
direction of the vectorial field solution.
\section{Non Linear Vectorial Lattice }
We now consider the full equation of motion, including
nonlinear on-site potential terms.
Extra nonlinear terms depending on
time derivatives may also appear in the case of non Cartesian coordinates systems:
\begin{eqnarray}
\label{emnl}
\frac{\partial^2 E(n,\alpha,t)}{\partial t^2 } =
- \sum_{n',\alpha'} J (n-n',\alpha,\alpha') E(n',\alpha',t)+ \nonumber\\
\sum_{d=0}^{2} \sum_{k=0}^{d}\sum_{\alpha'',\alpha'\leq \alpha''}
c_{d,k}^{\alpha}(\alpha',\alpha'') E^{(k)}(n,\alpha',t)E^{(d-k)}(n,\alpha'',t)
+ \nonumber\\
\sum_{d=0}^{2} \sum_{k=0}^{d}\sum_{j=0}^{k}
\sum_{\alpha''',\alpha''\leq \alpha''', \alpha'\leq \alpha'' } C_{d,k,j}^{\alpha}(\alpha',\alpha'',\alpha''')
E^{(j)}(n,\alpha',t)E^{(k-j)}(n,\alpha'',t)E^{(d-k)}(n,\alpha''',t)
\end{eqnarray}
where $E^{(j)}(n,\alpha,t)$ indicates the $j-$th time derivative of
$E(n,\alpha,t)$, and $c_{d,k}^{\alpha}(\alpha',\alpha'')$,
$C_{d,k,j}^{\alpha}(\alpha',\alpha'',\alpha''')$ are the quadratic
and cubic nonlinear terms numerical coefficients. Index $d$ is the total
time derivative order of each term. The terms with $d=0$ are the
nonlinear potential force terms; the others derive from the kinetic
energy in the case of non Cartesian coordinates so that $d \leq 2$.
The quadratic terms in (\ref{emnl}) give rise to second harmonic and
constant terms that, if we
look for small amplitude solution, have to be included as
additional smaller corrections:
\begin{eqnarray}
\label{wpc}
\vec{E}_l(n,t) &=& \epsilon \, e^{i(q_0n_0-\omega_l t_0)}\,
(\vec{V}_{l}-i \epsilon \vec {V^{(1)}_l} \frac{\partial }{\partial x_1})
A(x_1,t_1,t_2) \nonumber \\
&+& \epsilon^2 \,
e^{2i(q_0n_0-\omega_l t_0)}\,
\vec\gamma_l \, A^2(x_1,t_1,t_2)
+\epsilon^2 \, \vec\mu_l \, |A(x_1,t_1,t_2)|^2 + O(\epsilon^3) \,.
\end{eqnarray}
We are interested in situations where dispersion can balance
nonlinearity, and therefore have to be measured by the same scaling
parameter $\epsilon$. While the overall $\epsilon$ factor was not
important in the linear case, it must be explicitly included in the
nonlinear case.
We solve the equation of motion (\ref{emnl}) on the three
characteristic magnitude scales of the wave
packet. At order $\epsilon$ we get
$\omega_l$ ,$\vec{V}_{l}$ from equation (\ref{o1}).
At order $\epsilon^2$ we get for the wave packet term
expressed in the form (\ref{wpuno})
the
system of equations (\ref{o2}) for each $q_1$ in the integration
range. After integration
on the envelope distribution this gives rise to the equation in the
anti-transformed Fourier space
\begin{equation}
\label{oe}
\left( 2\omega_l \vec{V}_l \frac{\partial }{\partial t_1} +
(\hat{J} - \omega^2_l) \vec{V}^{(1)}_l \frac{\partial }{\partial x_1}
+ \hat{J}' \vec{V}_l \frac{\partial }{\partial x_1} \right)
A(x_1,t_1,t_2)=0 \; .
\end{equation}
In fact from (\ref{ap1}), the average wave numbers deviation is
$ <q_1>= -i \partial A / \partial x_1$ and the averaged frequency deviation
is in the same way
$<\Delta \omega_l>=<\omega_l^{(1)}q_1>= i \partial A / \partial t_1$.
By scalar product of (\ref{oe}) and $V^*_m$ $ \forall m\neq l$ we obtain the
components $\alpha_m$ (\ref{am}) of $\vec{V}^{(1)}_l$ (\ref{vp}) on the
base $\{\vec{V}_m\}$ and by scalar product of (\ref{oe}) and $V^*_l$ we obtain
the equation (\ref{wp1}) with $\omega^{(1)}_l$ defined by (\ref{op}).
At the same order of expansion one determines the vectors
$\vec{\gamma}_l$, $\vec{\mu}_l$ collecting the terms of
corresponding order $(\epsilon^{2})$ and phase in the equation of
motion (\ref{emnl}) in which the solution form (\ref{wpc})
has been inserted \cite{nota}. One then obtains
$\vec{\gamma}_l$ by solving the algebraic system
\begin{equation}
\label{ga}
(\hat{J}(2q_0)-4 \omega^2_l(q_0))\vec{\gamma}_l=
\sum_{d=0}^{2}\sum_{k=0}^{d} \sum_{\alpha'',\alpha'\leq \alpha''}
\vec{c}_{d,k}(\alpha',\alpha'') (-i\omega_l)^d V_l(\alpha')V_l(\alpha'')
\end{equation}
where $\vec{c}_{d,k}(\alpha',\alpha'')$ is the vector of components
${c}^{\alpha}_{d,k}(\alpha',\alpha'')$, each derivative with respect
to $t_0$ giving a factor $(-i\omega_l)$.
And $\vec{\mu}_l$ is obtained from the system
\begin{equation}
\label{mu}
\hat{J}(0)\vec{\mu}_l=
\sum_{d=0}^{2}\sum_{k=0}^{d} \sum_{\alpha'',\alpha'\leq \alpha''}
\vec{c}_{d,k}(\alpha',\alpha'')[(i\omega_l)^k (-i\omega_l)^{d-k}
V^*_l(\alpha')V_l(\alpha'')+(-i\omega_l)^k (i\omega_l)^{d-k}
V^*_l(\alpha'')V_l(\alpha')] \,.
\end{equation}
At order $\epsilon^3$, from (\ref{emnl}) and (\ref{wpc})
for the terms in $ e^{i(q_0n_0-\omega_l(q_0)t_0)}$, one obtains
the system of equations:
\begin{eqnarray}
\label{2r}
\left(
[(\frac{1}{2} \hat{J}''-{\omega^{(1)}_l}^2)\vec{V}_l+
(\hat{J}'-2\omega_l\omega^{(1)}_l)\vec{V}^{(1)}_l+
(\frac{1}{2}(\hat{J}-\omega^2_l) \vec{V}^{(2)})]
\frac{\partial^2}{\partial s_1^2}+
2 i\omega_l \vec{V}_l \frac{\partial}{\partial t_2}\right)\,
A(s_1,t_2)+ \nonumber \\
\vec{\cal{Q}} |A(s_1,t_2)|^2 A(s_1,t_2) =0
\end{eqnarray}
where
\begin{eqnarray}
\label{calq}
\vec{\cal{Q}}=
\sum_{d=0}^{2}\sum_{k=0}^{d} \sum_{\alpha'',\alpha'\leq \alpha''}
\vec{c}_{d,k}(\alpha',\alpha'')[ (i\omega_l)^k (-2i\omega_l)^{d-k}
V^*_l(\alpha')\gamma_l(\alpha'')+(-2i\omega_l)^k (i\omega_l)^{d-k}
V^*_l(\alpha'')\gamma_l(\alpha')]+ \nonumber\\
2\sum_{d=0}^{2}\sum_{\alpha'',\alpha'\leq \alpha''}
\vec{c}_{d,0}(\alpha',\alpha'')(-i\omega_l)^d \mu_l(\alpha')V_l(\alpha'')
+\nonumber \\
\sum_{d=0}^{2} \sum_{k=0}^{d}\sum_{j=0}^{k}
\sum_{\alpha''',\alpha''\leq \alpha''', \alpha'\leq \alpha'' }
\vec{C}_{d,k,j}(\alpha',\alpha'',\alpha''')
[(-i\omega_l)^j(-i\omega_l)^{k-j} (i\omega_l)^{d-k}
V_l(\alpha')V_l(\alpha'')V_l^*(\alpha''')+ \nonumber \\
(-i\omega_l)^j(i\omega_l)^{k-j} (-i\omega_l)^{d-k}
V_l(\alpha')V^*_l(\alpha'')V_l(\alpha''')+
(i\omega_l)^j(-i\omega_l)^{k-j} (-i\omega_l)^{d-k}
V^*_l(\alpha')V_l(\alpha'')V_l(\alpha''')]
\end{eqnarray}
The first line of the equation (\ref{2r}) corresponds to the third
order expansion (\ref{o3}) of the linear operator equation
(\ref{d}) applied to the wave packet (terms in
$e^{i(q_0n_0-\omega_l(q_0)t_0)}$ in (\ref{wpc})), in the moving
reference frame used in (\ref{ls}),
$<q_1>= -i ({\partial
}/{\partial s_1}) A(s_1,t_2)$, and $<\omega^{(2)}_l {q_1^2}/{2}> =
<\Delta(\Delta(\omega_l))>= i (\partial / \partial t_2 ) A(s_1,t_2)$.
The first two lines of the non linear coefficients vector
$\vec{\cal{Q}}$ arise from the double product,
in the nonlinear quadratic force terms in (\ref{emnl}),
between $O(\epsilon)$ and $O(\epsilon^2)$ components of (\ref{wpc});
the last two lines come from the nonlinear cubic
force terms in (\ref{emnl}) when considering just the
$O(\epsilon)$ terms in $\vec{E}_l(n,t)$.
Multiplying Eq.~(\ref{2r}) by $\vec{V}_l^*$ we obtain the NLS equation:
\begin{equation}
\label{nls}
\left( P \frac{\partial^2}{\partial s_1^2}+
i \frac{\partial}{\partial t_2} \right)
A(s_1,t_2)+ Q |A(s_1,t_2)|^2 A(s_1,t_2) =0
\end{equation}
where
\begin{equation}
\label{p}
P=\frac{1}{2\omega_l}
\left( \vec{V}^*_l \frac{\hat{J}''}{2}\vec{V}_l
- (\frac{ {\vec{V}_l}^* \hat{J}' \vec{V}_l}{2 \omega_l})^2
+ \sum_{m\neq l} \frac{| {\vec{V}_m}^* \hat{J}' \vec{V}_l|^2}
{ \omega^2_l-\omega^2_m} \right)
\end{equation}
and
\begin{equation}
\label{q}
Q=\frac{\vec{V}^*_l \vec{\cal{Q}}}{2\omega_l}
\end{equation}
The first part of equation (\ref{nls}) is equation (\ref{ls}) for
the wave packet with $2P=\omega^{(2)}_l$ given by the (\ref{os}) to which
is now added the nonlinear part with coefficient $Q$.
If $PQ > 0$ then the effect of the amplitude dependent non linear
potential well in (\ref{nls}) balances the wave packet group velocity
dispersion giving rise to the stable envelope soliton solution \cite{rem}.
\section{Summary.}
The main outlines of the approach we introduce are the following.
The envelope soliton like solutions arise in systems
with weak dispersion and weak nonlinearity by two parallel
series expansions driven by a common expansion parameter ($\epsilon$):
on one hand the weakness of the diffusion, for a wave packet like solution,
allows an expansion of the equations that
regulate the space time behaviour of the solution
on different scales; on the other hand the weak nonlinearity, for small
amplitude solutions, allows to write the equations of motion at
increasing orders of accuracy introducing the nonlinear terms in a
progressive way.
For scalar fields the Taylor series expansion of dispersion relations
gives directly the diffusive part for the envelope equations of
motion. Vectorial fields are instead characterized by a linear part
which gives rise, in the $q$ space, to
an eigenvalue (dispersion relations) and
eigenvector (relative amplitude of the different components)
problem (\ref{d}) ;
to obtain the correct expansion in multiple scales it is then
necessary to apply the perturbation theory (\ref{v}),(\ref{omega}).
Finally, to combine this perturbative expansion with the
nonlinear one, we antitransform the equations (\ref{o2}), (\ref{o3})
as done in (\ref{oe}), (\ref{2r}).
Following the approach introduced in this paper it is straightforward
to derive the NLS equation for every non linear vectorial lattice
with on-site nonlinearities and with an arbitrary number of
components. For more complex systems this could even be programmed in symbolic
languages to provide a fully automatic method.
After having identified the nonlinear coefficients
$c_{d,k}^{\alpha}(\alpha',\alpha'')$ and
$C_{d,k,j}^{\alpha}(\alpha',\alpha'',\alpha''')$ in the equation of
motion (\ref{emnl}), there are only algebraic systems to solve:
one has to solve the eigenvectors $V_l(q_0)$ and
the eigenvalues $\omega_l(q_0)$ of the matrix $\hat{J}(q_0)$,
then the systems (\ref{ga}) and (\ref{mu}) for $\gamma_l$ and
$\mu_l$, and to derive
$P$ from (\ref{p}) and $Q$ from
(\ref{calq}) and (\ref{q}).
{From} (\ref{nls}) one then obtains, if $PQ \geq 0$, the envelope
function $A(x_1,t_1,t_2)$ that, inserted into (\ref{wpc})
together with the eigenvector correction (\ref{vp}), (\ref{am}),
gives the complete $O(\epsilon^2)$ solution we are looking for.
\section*{Acknowledgments}
We thank R. Monasson for a critical reading of the manuscript.
|
\section{Introduction}
Over the past few years our understanding of string theory has
developed considerably. We now know that the five superstring
theories as well as low-energy 11-dimensional supergravity are related
through an intricate series of dualities and it has been argued that
all these theories are limits of an underlying 11-dimensional theory
called ``M-theory'' \cite{Witten-various} whose microscopic
description is as yet unknown. It has been found that in addition to
one-dimensional stringlike excitations there are higher-dimensional
branes in each of these theories which may in some regimes be
considered to be just as fundamental as the strings. In the five
superstring theories there are D-branes of various dimensions
\cite{Polchinski} as well as the fundamental string and NS5-branes.
In M-theory there are M2-branes and M5-branes which are related to the
branes of the superstring theories through various duality
transformations.
A fundamental class of problems is the identification of the
world-volume action for the various branes appearing in the six
theories of interest. This problem can be posed in a number of
contexts of differing degrees of complexity. The simplest problem is
to find the low-energy action for a single brane in a flat background
metric with no nontrivial background supergravity fields. A more
difficult problem is to find the action for a single brane in an
arbitrary background metric and field configuration which satisfies
the supergravity equations of motion. The problem can be made still
harder by considering systems of many branes, either in a flat or
general background. Even for a single fundamental superstring the
action in a general background including arbitrary R-R fields is not
yet well understood; for recent work in this direction see, for
example, \cite{bvw} and references therein. For single D-branes the
situation is somewhat better. The action for a single D-brane moving
in a general background is the Born-Infeld action \cite{Leigh}, which
reduces to the maximally supersymmetric $U(1)$ super Yang-Mills theory
on the world volume in the case where the brane is almost flat and has
only low-energy excitations. This action is supplemented by
Chern-Simons type couplings to background R-R fields
\cite{Douglas,Li-bound}. Even for the single D-brane, however, there
are subtle issues involved in giving a world-volume supersymmetric
description of the Born-Infeld theory. For
systems of $N$ D-branes, it is known that the low-energy
action for parallel branes in a flat background is given by the
supersymmetric $U(N)$ super Yang-Mills theory found by dimensional
reduction from 10D \cite{Witten-bound}. So far there has been little
progress in describing the action governing systems of many D-branes
in a general background. This problem is due in part to the absence
of a nonabelian generalization of the Born-Infeld action (although one
proposal for such an action was made in \cite{Tseytlin}), and in part
to ordering problems which arise even in the low-energy theory in the
presence of general backgrounds.
In this paper we consider the simplest system of many D-branes in a
general background: low-energy configurations of many D0-branes moving
in an arbitrary but weak background of type IIA supergravity.
According to an argument of Seiberg \cite{Seiberg-DLCQ} (see also
\cite{Sen-DLCQ}), the action for such a system of
D0-branes should be related to the DLCQ description of M-theory in an
associated 11-dimensional supergravity background. This generalizes
the BFSS matrix theory conjecture \cite{BFSS,Susskind-DLCQ}, which
states that supersymmetric matrix quantum mechanics (the low-energy
theory of $N$ D0-branes in flat space) gives a light-front description
of M-theory in a flat background. In a previous paper
\cite{Mark-Wati-3} we used a matrix theory formulation of the
multipole moments of the supercurrent components in 11D supergravity
(derived in \cite{Mark-Wati,Dan-Wati-2,Mark-Wati-3}) to propose an
explicit description of the matrix theory action up to terms linear in
the background fields, as well as an algorithm for using higher-loop
calculations in matrix theory to find the higher order terms in the
matrix theory action in general backgrounds. In this paper, we use
our proposal for the general background Matrix theory action and
follow the arguments of Seiberg to deduce the leading terms in the
action
for multiple D0-branes in weak type IIA supergravity backgrounds. We
then perform some simple tests of the
Matrix theory action and the related multiple D0-brane action. In the
D0-brane case, we show that our prescription satisfies a constraint
originally suggested by Douglas \cite{Douglas-curved} that the masses
of off-diagonal matrix elements between a pair of separated D0-branes
agree with the minimal geodesic length between the D0-branes. This
property holds also in the Matrix theory case where the separated
objects are a pair of gravitons, and we use it to show that the
leading order potential between a pair of gravitons in a weakly curved
Ricci-flat background is correctly reproduced by the proposed general
background
Matrix theory action.
The paper is organized as follows. In Section 2 we review our
proposal for the linear terms in the general background Matrix theory
action. Then, using this action, we follow the arguments of Seiberg
to deduce leading terms in the action for multiple D0 branes in an
arbitrary weak type IIA supergravity background. In section 3 we
describe tests of the IIA and matrix theory actions. We conclude in
section 4 with a discussion of related issues and comments on further
directions.
\section{Linear coupling to backgrounds}
In subsection \ref{sec:matrix-background} we recall the proposal made
in \cite{Mark-Wati-3} for the terms in the action of matrix theory
which are linear in the background fields. In subsection
\ref{sec:0-backgrounds} we use the approach of Seiberg to relate this
matrix theory action to an action for multiple D0-branes in IIA
background fields. This allows us to deduce the leading
terms in the multiple D0-brane action, which to the best of our
knowledge have not been previously described.
\subsection{Backgrounds in matrix theory}
\label{sec:matrix-background}
In \cite{Mark-Wati-3} we proposed that the linear effects of a general
matrix theory background with metric $g_{IJ} = \eta_{IJ} + h_{IJ}$ and
3-form field $A_{IJK}$ could be described by supplementing the flat
space matrix theory action
\begin{equation}
S_{{\rm flat}} = -{1 \over 2 R} \int d\tau \; {\rm Tr} \, \biggl\lbrace
-D_{\tau} X_i D_{\tau} X_i + \frac{1}{2} [X_i,X_j] [X_i,X_j] +
\Theta_{\alpha} D_{\tau} \Theta_{\alpha} - \Theta_{\alpha}
\gamma^i_{\alpha \beta} [X_i,\Theta_{\beta}] \biggr\rbrace
\label{eq:flat-action}
\end{equation}
with
additional terms of the form
\begin{eqnarray}
S_{\rm weak} & = & \int d \tau \;
\sum_{n = 0}^{\infty} \sum_{i_1, \ldots, i_n}\frac{1}{n!}
\left(
\frac{1}{2}
T^{IJ (i_1 \cdots i_n)} \partial_{i_1} \cdots \partial_{i_n} h_{IJ}
(0)
+
J^{IJK (i_1 \cdots
i_n)}
\partial_{i_1} \cdots \partial_{i_n} A_{IJK} (0)
\right. \nonumber \\
& &\hspace{1.2in} \left.
+M^{IJKLMN (i_1
\cdots i_n)} \partial_{i_1} \cdots \partial_{i_n} A^D_{IJKLMN} (0)
+ {\rm fermion \; terms} \right) \label{eq:general-background}
\end{eqnarray}
where $A^D$ is the dual 6-form field which satisfies at linear order
\[
dA^D ={}^* (dA).
\]
In (\ref{eq:general-background}) the matrix expressions $T^{IJ (i_1
\cdots i_n)}, J^{IJK (i_1 \cdots
i_n)}, M^{IJKLMN (i_1
\cdots i_n)} $ are the matrix theory forms of the multipole moments of
the stress-energy tensor, membrane current and 5-brane current of 11D
supergravity. Explicit forms for the parts of these moments depending
only on the 9 bosonic transverse matrices $X^i$
were given in \cite{Dan-Wati-2}, and the terms quadratic in the
fermions were given in \cite{Mark-Wati-3}, as well as some terms
quartic in the fermions. For example, the zeroeth moments of the
components of the stress-energy tensor are given by
\begin{eqnarray}
T^{++} &=& {1 \over R}{\rm STr} \,\left({\rlap{1} \hskip 1.6pt \hbox{1}}\right)\nonumber\\
T^{+i} &=& {1 \over R}{\rm STr} \,\left(D_t X_i\right)\nonumber\\
T^{+-} &=& {1 \over R}{\rm STr} \,\left({1 \over 2} D_t X_i D_t X_i + {1
\over 4}
F_{ij}
F_{ij} + {1 \over 2} \Theta\gamma^i[X^i,\Theta]\right)\nonumber\\
T^{ij} &=& {1 \over R}{\rm STr} \,\left( D_t X_i D_t X_j + F_{ik} F_{kj}
- {1
\over 4}
\Theta\gamma^i[X_j,\Theta] - {1 \over 4}
\Theta\gamma^j[X_i,\Theta]\right)\nonumber\\
T^{-i} &=& {1 \over R} {\rm STr} \,\left({1 \over
2}D_t X_iD_t X_jD_t X_j + {1 \over 4} D_t X_i F_{jk} F_{jk} +
F_{ij} F_{jk} D_t X_k\right) \label{eq:stress-tensor}\\ & & - {1
\over 4R} {\rm STr} \,\left(\Theta_\alpha
D_t X_k[X_m,\Theta_\beta]\right)\{\gamma^k\delta_{im}
+\gamma^i\delta_{mk} -2\gamma^m\delta_{ki} \}_{\alpha
\beta}\nonumber\\ & & - {1 \over 8R} {\rm STr} \,\left(\Theta_{\alpha}
F_{kl}[X_m,\Theta_{\beta}]\right)\{ \gamma^{[iklm]} + 2 \gamma^{[lm]}
\delta_{ki} + 4\delta_{ki}\delta_{lm} \}_{\alpha \beta}\nonumber\\ & &
+ {i \over 8R} {\rm Tr} \,(\Theta \gamma^{[ki]} \Theta \; \Theta \gamma^k
\Theta)\nonumber\\
T_f^{--} &=& {1 \over 4R} {\rm STr} \,\left(F_{ab}F_{bc}F_{cd}F_{da} - {1 \over
4}F_{ab}
F_{ab} F_{cd} F_{cd} + {\Theta} \Gamma^b \Gamma^c \Gamma^d F_{ab}
F_{cd}
D_a\Theta + {\cal O} ({\Theta^4})\right)\nonumber
\end{eqnarray}
where ${\rm STr}$ indicates a trace which is symmetrized over all
orderings of terms of the forms $F_{ab}, \Theta$ and $[X^i, \Theta]$,
indices $i (a)$ run from 1 (0) through 9, and we have defined
$F_{0i} = D_t{X}^i,
F_{ij} = i[X^i, X^j]$. There are two types of terms which contribute
to higher moments of these components of the stress-energy tensor
\begin{equation}
T^{IJ (i_1 i_2 \cdots i_n)} =
{\rm Sym} \, (T^{IJ}; X^{i_1}, X^{i_2}, \ldots, X^{i_n}) +
T_{\rm fermion}^{IJ(i_1 i_2 \cdots i_n)}
\end{equation}
The contributions ${\rm Sym} \, ({\rm STr}\; (Y); X^{i_1}, \ldots, X^{i_n})$
are defined as the symmetrized average over all possible orderings
when the matrices $X^{i_k}$ are inserted into the trace of any product
$Y$ of the forms $F_{ab}, \Theta, [X^i, \Theta]$. In general there
are additional fermionic contributions of arbitrary order to the
higher multipole moments, of which the simplest example is the spin
contribution to the angular momentum
\begin{equation}
T_{\rm fermion}^{+i(j)} = {1 \over 8R} {\rm Tr} \,(\Theta \gamma^{[ij]}
\Theta)
\end{equation}
The precise form of these fermionic contributions will not be
important to us in this paper, for reasons which will be discussed in
section \ref{sec:oscillator-masses}.
The results of \cite{Mark-Wati-3} for the matrix membrane and 5-brane
currents are reproduced in the Appendix for convenience. With these
definitions, (\ref{eq:general-background}) gives a formulation of
matrix theory in a weak background metric to first order in the metric
$h_{IJ}$, the 3-form $A_{IJK}$, and all their higher derivatives. It
was argued in \cite{Mark-Wati-3} that if the matrix theory conjecture
is true in flat space, this formulation must be correct at least to
order $\partial^4h, \partial^4A$ for a class of backgrounds which can
be produced as the long range fields around supergravity sources
described by matrix theory objects. We conjectured further that this
form may work to all orders and in a general background. It should be
emphasized, however, that this formulation can only be given for
M-theory backgrounds with a global $U(1)$ symmetry around a compact
direction, as we do not know how to incorporate dependence of the
background on the compact coordinate $x^-$. We only expect this
action to be part of a consistent all-orders matrix theory action in a
general background when the background satisfies the equations of
motion. The derivation of this action also depended upon a particular
choice of gauge for the graviton, so that it may be necessary to
restrict attention to background fields satisfying the linearized
gauge
\begin{equation}
\partial^I \bar{h}_{IJ} =\partial^I (h_{IJ}-\frac{1}{2}
\eta_{IJ}h_{K}^{\; K}) = 0.
\end{equation}
\subsection{Backgrounds for D0-branes}
\label{sec:0-backgrounds}
We now investigate how the Matrix theory action described in the
previous
section is related to the action for multiple D0-branes in background
type IIA supergravity fields.
In the case of a flat background, the
Matrix theory action may be derived by showing an equivalence between
the DLCQ limit of M-theory in a flat background with N units of momentum
around the circle and a particular limit of type IIA string theory with
N D0-branes \cite{Seiberg-DLCQ}. In this limit, the only remaining
degrees of freedom are the lowest energy modes of open strings ending on
the N D0-branes. The dynamics of these modes are in general described
by a non-abelian generalization of the Born-Infeld action whose complete
form is not known. However, in the appropriate limit of type IIA string
theory, most of the terms in this action vanish, and we find that the
dynamics of DLCQ M-theory in a flat background are described by an
action equivalent to the dimensional reduction of D=10 super Yang-Mills
theory to 0+1 dimensions.
The action for Matrix theory with background supergravity fields
given in the previous section has been derived completely within the
context of Matrix theory. However, in principle, one should be able to
apply Seiberg's arguments to this case also and derive the same action
as a limit of the action for D0-branes in type IIA string theory with
background supergravity fields. Again, only particular terms in the
D-brane action will survive in the appropriate limit, but unlike the
flat space case, not even these terms are known except in the case of a
single brane. Hence, in the case $N=1$, we should be able to rederive
our result from previously known facts about D-branes, but more
importantly, we will be able to apply the arguments in reverse for $N>1$
to derive previously unknown leading terms in the action for multiple
D0-branes in an arbitrary weak type IIA supergravity background. Using
T-duality, our result may be extended to give leading terms in the
actions for all other types of D-branes.
\subsubsection{Relationship between DLCQ and type IIA backgrounds}
We now review the steps taken in \cite{Seiberg-DLCQ} as they apply in
the case of weak backgrounds to make precise the relationship between
the matrix theory action and the multiple D0-brane actions. In
particular, we must determine the relationship between the D=11
supergravity fields appearing in the Matrix theory action
(\ref{eq:general-background})
and the
type IIA supergravity fields appearing in the related
D0-brane action.
We start by considering M-theory with background metric
\[
{g}_{IJ} = \eta_{IJ} + {h}_{IJ}
\]
in a frame with a compact coordinate $x^-$ of size $R$ which is
light-like in the flat space limit $h_{IJ} = 0$. This theory can be
described as a limit of a family of space-like compactified theories.
Defining an $\hat{M}$-theory with background metric
\[
\hat{g}_{IJ} = \eta_{IJ} + \hat{h}_{IJ}
\]
in a frame with a spacelike compact direction $x^{10}$ of size $R_s$,
the DLCQ limit of the original M-theory can be found by boosting the
$\hat{M}$-theory in the $x^{10}$ direction with boost parameter
\[
\gamma = \sqrt{{R^2 \over 2 R_s^2} + 1}
\]
and then taking a limit $R_s \rightarrow 0$.
The metric $\hat{g}_{IJ}$ in the $\hat{M}$-theory
is related to that of the original M-theory by
\begin{eqnarray*}
\hat{h}_{ij} &=& h_{ij}\\
\hat{h}_{0\,i} &=& {\alpha \over \sqrt{2}}h_{+i} + {1 \over
\alpha\sqrt{2}} h_{-i}\\
\hat{h}_{10\,i} &=& {\alpha \over \sqrt{2}}h_{+i} - {1 \over
\alpha\sqrt{2}} h_{-i}\\
\hat{h}_{0\,0} &=& h_{+-} + {\alpha^2 \over 2}h_{++} + {1 \over
2 \alpha^2}h_{--}\\
\hat{h}_{10 \, 10} &=& -h_{+-} + {\alpha^2 \over 2}h_{++} + {1 \over
2 \alpha^2}h_{--}\\
\hat{h}_{0\,10} &=& {\alpha^2 \over 2}h_{++} - {1 \over
2 \alpha^2}h_{--}\\
\end{eqnarray*}
where we have defined
\begin{eqnarray*}
\alpha &=& \gamma(1-v) = \gamma - \sqrt{\gamma^2 - 1}\\
&=& {R_s \over R \sqrt{2}} + {\cal O}((R_s/R)^3)
\end{eqnarray*}
M-theory on a small spacelike circle of radius $R_s$
is equivalent to type IIA string theory with background fields given to
leading order by:
\begin{eqnarray*}
h^{IIA}_{\mu \nu} &=& \hat{h}_{\mu \nu} + {1 \over 2} \eta_{\mu \nu}
\hat{h}_{10\,10}\\
C_{\mu} &=& \hat{h}_{10 \, \mu}\\
\phi &=& {3 \over 4} \hat{h}_{10 \, 10}
\end{eqnarray*}
and parameters
\[
g_s = (R_s M_p)^{3/2}, \; \; \; M_s = R_s^{1/2}M_p^{3/2}
\]
where $M_p$ is the eleven-dimensional Planck mass. Here we have defined
$g_s$ to be a constant and chosen the dilaton $\phi$ so that $\phi=0$ in
the case of a circle of constant size $R_s$ with $h_{10 \; 10} = 0$.
Thus the effective string coupling is given locally by the combination
\[
g_s(\vec{x}) = g_s e^{\phi} .
\]
\\
Combining the two equivalences, we conclude that DLCQ M-theory with N
units of momentum on a lightlike circle of size $R$ and background
metric $g_{IJ}$ is equivalent to the $R_s \rightarrow 0$ limit of type
IIA string theory with N D0-branes, parameters
\[
g_s = (R_s M_p)^{3/2}, \; \; \; M_s = R_s^{1/2}M_p^{3/2}
\]
and background fields
\begin{eqnarray}
h^{IIA}_{0 0} &=& {3 \over 2}h_{+-} + {\alpha^2 \over 4}h_{++} + {1
\over
4 \alpha^2} h_{--} \nonumber\\
h^{IIA}_{0 i} &=& {\alpha \over \sqrt{2}} h_{+i} + {1 \over \alpha
\sqrt{2}} h_{-i}
\nonumber\\
h^{IIA}_{i j} &=& h_{ij} + {1 \over 2} \delta_{ij} (-h_{+-} + {\alpha^2
\over 2}h_{++} + {1 \over 2 \alpha^2} h_{--}) \label{eq:relations}\\
\phi &=& -{3 \over 4}h_{+-} + {3 \alpha^2 \over 8}h_{++} + {3 \over
8 \alpha^2} h_{--}\nonumber\\
C_0 &=& {\alpha^2 \over 2}h_{++} - {1 \over 2 \alpha^2} h_{--}
\nonumber\\
C_i &=& {\alpha \over \sqrt{2}} h_{+i} - {1 \over \alpha \sqrt{2}}
h_{-i}
\nonumber
\end{eqnarray}
At first glance, such a limit seems problematic. In particular, it
appears that for fixed finite values of the DLCQ metric components,
the background fields of the equivalent type IIA theory diverge in the
limit $R_s \rightarrow 0$ since $1/\alpha \rightarrow \infty$. However,
recall from the flat space case
that without a further rescaling of the parameters in the type IIA
picture, the energies of the states we are interested in go to 0 like
$R_s$. As we shall see, the appropriate rescaling of parameters which
makes the energies we are interested in finite without changing the
physics also ensures that the apparent divergences of background field
components do not lead to divergent terms in the final action.
Another feature of this action is that after the appropriate
rescaling the characteristic length scale $L$ associated with the
structure of the metric becomes much smaller than the string length
$1/M_s$. While this may seem unusual, it is precisely what is needed
for the physics of the system to be completely captured by the open
string theory describing the D0-brane theory at substring scales
studied in \cite{DKPS}. Indeed, for compact manifolds such as tori,
it is this effect which makes it possible for the wrapped string modes
corresponding to momentum excitations on the dual space to
become physically relevant \cite{WT-compact,Seiberg-DLCQ}.
\subsubsection{$N=1$ actions}
We now use the correspondence just discussed to make an explicit
comparison between the matrix theory and IIA descriptions of a system
of $N$ 0-branes in a weak background field. We begin with the case
$N=1$. Here, both the Matrix theory and D0-brane actions are known,
so we would like to check that the Matrix theory action may be derived
from the D0-brane action before proceeding to the case $N>1$ where the
D0-brane action is not known. For the case $N=1$, the Matrix theory
action (\ref{eq:flat-action},\ref{eq:general-background}) reduces to
\begin{eqnarray}
S & = & {1 \over R} \int dt \left\{ \frac{\dot{x}^2}{2} + \frac{1}{2}
h_{++}(\vec{x}) +
h_{+i}(\vec{x}) \dot{x}^i
+ {1 \over 2} h_{ij}(\vec{x}) \dot{x}^i \dot{x}^j
\label{graviton} \right.\\
& &\hspace{1in} + \left.\frac{1}{2} h_{+-}(\vec{x})\dot{x}^2 +
\frac{1}{2}
h_{-i}(\vec{x}) \dot{x}^2 \dot{x}^i + \frac{1}{8} h_{--}(\vec{x})
\dot{x}^4 \right\}. \nonumber
\end{eqnarray}
In this case we expect the action to describe a single graviton in
curved space with unit momentum along the lightlike circle. Such an
action was derived from supergravity in \cite{Mark-Wati}; expression
(\ref{graviton}) is indeed identical to the supergravity result given by
equation (13) in that paper.
The world-volume action for a single D0-brane moving in a general
type IIA background supergravity fields is given by
\begin{equation}
S_{IIA} = - \tau_0\int d \xi e^{-\phi} \sqrt{g_{\mu \nu}
(d{x}^{\mu}/d \xi )
(d{x}^{\nu}/d \xi)} +
\tau_0 \int C_{\mu} dx^{\mu}
\label{eq:single-0}
\end{equation}
where $\phi$, $g_{\mu \nu}$, and $C_{\mu}$ are the background dilaton,
metric, and R-R one-form fields, and the parameter $\tau_0$
is the D0 mass, given by
\[
\tau_0 = {M_s \over g_s} .
\]
One can also consider background R-R three form $C_{\mu \nu \lambda}$
and
NS-NS antisymmetric tensor $B_{\mu \nu}$ fields,
but these do not couple to a
single zero-brane.
According to the equivalence presented in the previous section, the
Matrix theory action (\ref{graviton}) should arise from the D0-brane
action (\ref{eq:single-0}) by rewriting the type IIA background fields
in terms of the desired DLCQ supergravity background using the
relations (\ref{eq:relations}), then rescaling parameters and taking
the limit $R_s \rightarrow 0$. We will now verify this explicitly.
Choosing a gauge in which the coordinate time $x^0$ is identified with
the worldvolume time $\xi$ we first expand the D0-brane action to
leading order in the background fields, giving
\begin{equation}
S = -\tau_0 \int d\xi \left\{ (1-v^2)^{1/2} (1- \phi) - {1 \over 2}
(1-v^2)^{-1/2}(h^{IIA}_{00} + 2h_{0i}v^i + h_{ij}v^iv^j) - C_0 - C_iv^i
\right\}
\label{eq:expand-Born-Infeld}
\end{equation}
where $v^i \equiv \dot{x}^i$. We now write the IIA background fields
in terms of the background fields in the equivalent DLCQ M-theory
using (\ref{eq:relations}). Keeping only the leading term in $R_s/R$
for each of the components of the metric $h_{IJ}$, we find
\begin{eqnarray}
S & = & {1 \over R_s} \int d \xi \left\{ -1 + {1 \over 2}
\frac{R^2_s}{R^2}
h_{++}(\vec{x}) + \frac{R_s}{ R}
h_{+i}(\vec{x}) v^i\right. \label{eq:replaced}\\
& &\hspace{1in} \left.
+ {1 \over 2} h_{ij}(\vec{x}) v^i v^j + \frac{1}{2}
h_{+-}(\vec{x})v^2 + \frac{v^2}{2}+\frac{1}{2}
\frac{R}{R_s} h_{-i}(\vec{x}) v^2 v^i + {1 \over 8}
\frac{R^2}{R_s^2} h_{--}(\vec{x}) v^4 \right\}. \nonumber
\end{eqnarray}
Many of these terms seem to diverge in
the $R_s \rightarrow 0$ limit we are interested in. However, as
mentioned above, this scaling is deceptive, since we must rescale
parameters in the theory so that the energies of the states we are
interested in remain finite rather than going to zero in the limit.
Indeed, from the fact that the conjugate momentum has a leading term
of order $v/R_s$ it can be seen that all the terms in
(\ref{eq:replaced}) which are linear in the background contribute to
the Hamiltonian at order $R_s$. Thus, as we need for the Seiberg
limit, the energy of the states of interest scale as $R_s$.
We may now perform the rescaling of \cite{Seiberg-DLCQ} by replacing
\[
R \rightarrow ({R_s \over R})^{1/2}R , \;\;\;\;\; \, \, \vec{x}
\rightarrow ({R_s \over R})^{1/2}\vec{x}, \;\;\;\;\; \, \, h(\vec{x})
\rightarrow h(\vec{x}).
\]
Note that the change of variables in the second replacement combines
with
the rescaling of dimensionful coefficients in the expansion of $h$ to
leave $h(\vec{x})$ unchanged, as suggested by the final replacement.
With these redefinitions the action (\ref{eq:replaced}) becomes
\begin{eqnarray}
S & = & \int d \xi \left\{ -\frac{1}{R_s} + {1 \over R}\left(
{1 \over 2} h_{++}(\vec{x}) +
h_{+i}(\vec{x}) v^i+ {1 \over 2} h_{ij}(\vec{x}) v^i v^j\right.
\right.\nonumber\\
& &\left.\left.\hspace{1.5in}
+ \frac{1}{2}
h_{+-}(\vec{x})v^2 + \frac{v^2}{2} +\frac{1}{2}
h_{-i}(\vec{x}) v^2 v^i +
{1 \over 8} h_{--}(\vec{x}) v^4 \right)\right\}. \nonumber
\end{eqnarray}
The first term is divergent in the $R_s \rightarrow 0$ limit and arises
from the
BPS energy of the single 0-brane; this term also appears in the flat
space theory and is discounted in the matrix theory limit.
Dropping this term gives precisely the matrix theory action
described by (\ref{graviton}) in the $N = 1$ case. Thus, we have
shown that the known Born-Infeld action for a single D0-brane
correctly reproduces the matrix theory action in a weak background
when the proper limit is taken.
\subsubsection{$N>1$ actions}
We now turn to the case $N>1$. Here, the appropriate action for
multiple D0-branes is not known, but by requiring that it reproduces the
general background Matrix theory action in the Seiberg limit, we will be
able to deduce its leading terms.
We first write down the D0-brane action in terms of the unknown
quantities coupling to the background fields. We define quantities
$I_x$ coupling linearly to each of the background fields,
so that to leading order in the background fields, the action for $N$ D0
branes is
\begin{eqnarray}
S &= & S_{{\rm flat}} +
\int dt \sum_{n=0}^{\infty} {1
\over n!} \left[
\frac{1}{2}
(\partial_{k_1}\cdots
\partial_{k_n} h^{IIA}_{\mu \nu}) \; I_h^{\mu \nu (k_1 \cdots k_n)}
+ (\partial_{k_1}\cdots \partial_{k_n} \phi)
\; I_{\phi}^{(k_1 \cdots k_n)}\right.\label{eq:IIA-general}\\
& & \hspace{1in}+ (\partial_{k_1}\cdots
\partial_{k_n} C_{\mu }) \; I_0^{\mu (k_1 \cdots k_n)}
+
(\partial_{k_1}\cdots
\partial_{k_n} \tilde{C}_{\mu \nu \lambda \rho \sigma \tau \zeta })
\; I_6^{\mu \nu \lambda
\rho \sigma \tau \zeta (k_1 \cdots k_n)}
\nonumber\\
& & \hspace{1in}+
(\partial_{k_1}\cdots \partial_{k_n}
B_{\mu \nu}) \; I_s^{\mu \nu (k_1 \cdots k_n)}
+
(\partial_{k_1}\cdots \partial_{k_n}
\tilde{B}_{\mu \nu \lambda \rho \sigma \tau}) \; I_5^{\mu \nu \lambda
\rho \sigma \tau
(k_1 \cdots k_n)} \nonumber\\
& &\hspace{1in} \left.+
(\partial_{k_1}\cdots
\partial_{k_n} C^{(3)}_{\mu \nu \lambda }) \; I_2^{\mu \nu \lambda (k_1
\cdots
k_n)}
+
(\partial_{k_1}\cdots
\partial_{k_n} \tilde{C}^{(3)}_{\mu \nu \lambda \rho \sigma })
\; I_4^{\mu \nu \lambda \rho \sigma (k_1 \cdots
k_n)}\right] \nonumber
\end{eqnarray}
Here, $S_{{\rm flat}}$ is the flat space action for N D0-branes,
whose leading
terms are the dimensional reduction of D=10 SYM theory to 0+1
dimensions. The complete form of the higher order terms in the flat
space action is not known,
but these terms vanish in the Matrix theory limit.
We assume that the background satisfies the source-free IIA
supergravity equations of motion so that the dual fields $\tilde{C},
\tilde{B}, \tilde{C}^{(3)}$ are well-defined 7-, 6- and 5-form fields
given (at linear order) by
\begin{equation}
d \tilde{C} ={}^* dC, \;\;\;\;\;
d \tilde{B} ={}^* dB, \;\;\;\;\;
d \tilde{C}^{(3)} ={}^* dC^{(3)}.
\end{equation}
The sources $I_{2n}$ are associated with Dirichlet $2n-$brane
currents, while the sources $I_{s}$ and $I_{5}$ are associated with
fundamental string and NS5-brane currents respectively. It may seem
surprising that a system of D0-branes can give rise to a nonzero
D2-brane, D4-brane or D6-brane charge. Indeed, the integrated higher
brane
charges must vanish for a system containing a finite number $N$ of
D0-branes. Even for $N = 2$, however, a D0-brane configuration can
have nonvanishing multipole moments of higher D-brane charges. This
essentially arises as the T-dual of the result that the $n$th power of
the
curvature form $F$ on a Dirichlet D$p$-brane carries $(p-2n)$-brane
charge \cite{Witten-small,Douglas}; see \cite{WT-Trieste} and
references therein for a further discussion of this issue.
The problem we address in this subsection is the determination of the
IIA currents $I_x$ under the assumption that this action
reproduces the matrix theory action (\ref{eq:general-background}) in
the Seiberg limit. As we will see, the leading terms in all the
currents other than $I_5$ can be determined and are related to the
matrix theory supercurrent components tabulated in the Appendix.
For the case $N=1$
we see from (\ref{eq:expand-Born-Infeld}) that the nonvanishing
source components $I_x$ are
\begin{eqnarray}
I_h^{00(k_1\cdots k_n)} &=& {1 \over R_s}(1-\dot{x}^2)^{-1/2} x^{k_1}
\cdots
x^{k_n} \nonumber\\
I_h^{0i(k_1\cdots k_n)} &=& {1 \over R_s}(1-\dot{x}^2)^{-1/2} \dot{x}^i
x^{k_1}
\cdots x^{k_n} \nonumber\\
I_h^{ij(k_1\cdots k_n)} &=& {1 \over R_s}(1-\dot{x}^2)^{-1/2} \dot{x}^i
\dot{x}^j
x^{k_1} \cdots x^{k_n} \label{eq:n1}\\
I_\phi^{(k_1\cdots k_n)} &=& {1 \over R_s}(1-\dot{x}^2)^{1/2} x^{k_1}
\cdots
x^{k_n}
\nonumber\\
I_0^{0(k_1\cdots k_n)} &=& {1 \over R_s} x^{k_1}
\cdots
x^{k_n} \nonumber\\
I_0^{i(k_1\cdots k_n)} &=& {1 \over R_s}\dot{x}^i x^{k_1}
\cdots
x^{k_n} \nonumber
\end{eqnarray}
In the nonabelian case $N > 1$, the quantities $I_x$ will be some
complicated
functions of the $N \times N$ hermitian matrices $X^i$ as well as the
fermionic matrices $\Theta$. For each $I$, we can make an expansion
analogous to expanding in velocities for the $N=1$ case. We write
\[
I_x = \sum I_{x[n]}
\]
where $n$ counts the dimension of a function of the matrices $X,
\Theta$, giving $X$ dimension 1, $\dot{X}$ dimension 2 and $\Theta$
dimension 3/2. If we do a similar expansion for the flat
space action $S_0$, we find that it is precisely the $n=4$ terms
that remain in the Matrix theory limit, the higher order terms being
scaled to zero. In the general background case, the Matrix theory
action will arise from terms $I_n$ with $n \le 8$ (and their higher
moments), so it is these
terms that our analysis will determine.
We now proceed just as in the $N=1$ case.
We begin by working through the details of the analysis for those terms
coupling to the IIA graviton, dilaton and R-R 1-form field. These
terms are the most complicated. The analysis for the remaining bosonic
fields
is described at the end of this section.
By the Seiberg equivalences, the Matrix theory action with background
supergravity fields should result from replacing the IIA background
fields in (\ref{eq:IIA-general}) with their DLCQ counterparts
(\ref{eq:relations}), rescaling parameters as above, and taking the
limit $R_s \rightarrow 0$. Before rescaling, we find that the D0-brane
action becomes
\begin{eqnarray}
S &= & S_{\rm flat} - \frac{1}{2}\int dt \left[ (\sqrt{2} \alpha)^{-2}
h_{--}
\{ {1 \over 2} I_h^{00} + {1 \over 2} I_h^{ii} + {3 \over 2} I_\phi -
2 I_0^0 \} \right.
\nonumber\\
& &\hspace{1in} + (\sqrt{2} \alpha)^{-1} h_{-i} \{2 I_h^{0i} - 2 I_0^i
\}\nonumber\\
& &\hspace{1in} + h_{ij} \{ I_h^{ij} \}\nonumber\\
& &\hspace{1in} + h_{+-} \{ {3 \over 2} I_h^{00} - {1 \over 2}
I_h^{ii} -
{3 \over 2}
I_\phi
\}\label{eq:IIA-rewritten}\\
& &\hspace{1in} + (\sqrt{2} \alpha) h_{+i} \{ I_h^{0i} + I_0^i
\}\nonumber\\
& &\hspace{1in} + (\sqrt{2} \alpha)^{2} h_{++} \{ {1 \over 8} I_h^{00} +
{1
\over
8}
I_h^{ii} + {3 \over 8} I_\phi + {1 \over 2} I_0^0 \} \nonumber\\
& &\hspace{1in} + \left. \{ {\rm higher \; moment \; terms} \}\right]
\nonumber
\end{eqnarray}
The higher moment terms have exactly the same form as the terms
written, for example the full set of terms linear in $h_{-i}$ would be
\[
\sum_{n=0}^{\infty} {R \over R_s} \partial_{k_1} \cdots \partial_{k_n}
h_{-i} \{2 I_h^{0i(k_1 \cdots k_n)} - 2I_0^{i(k_1 \cdots k_n)} \}
\]
Because the distance scale associated with the metric is rescaled
along with the transverse coordinates in the rescaling of
\cite{Seiberg-DLCQ}, the rescaling of the partial derivatives in these
expressions cancel the scaling of the moment indices. Thus,
the higher moment terms which remain in the $R_s
\rightarrow 0$ limit are precisely those corresponding to 0th moments
which remain in the limit. The only terms in (\ref{eq:IIA-rewritten})
which remain in the limit other than the leading divergent D0-brane
energy term should be finite terms corresponding to the matrix theory
action (\ref{eq:general-background}). All terms which are linear in
the background and carry positive powers of $R/R_s$ in the limit must
cancel for the IIA action to agree with the matrix theory action.
This gives a number of restrictions on the parts of the IIA currents
with particular scaling dimensions. The constraints arising in this
fashion for the integrated (monopole) currents are
\junk{
In order to see what happens to this action under the appropriate
rescaling and change of variables, we must understand how the various
terms $I_[n]$ are affected. In the $N=1$ case considered above, each
$I_[n]$
consists of a single term and it is easy to see that the general rule is
\begin{equation}
(\partial_{k_1} \cdots \partial_{k_l} h) I_n \rightarrow ({R_s \over
R})^{(l+n)/2} (\partial_{k_1} \cdots \partial_{k_l} h) I_n
\label{scaling}
\end{equation}
For $N>1$, the $I_n$ will involve multiple terms involving both bosonic
and fermionic fields. However, in order that the flat space Matrix
theory arise correctly from the terms $(L_{flat})_2$ in the D0-brane
action, it must be that all terms in a given $I_n$ transform in the same
way so that the relations (\ref{scaling}) are still valid.
The expansion coefficients of the metric
\[
\partial_{k_1} \cdots \partial_{k_l} h
\]
have mass dimension $l$, so they are multiplied by $(R/R_s)^{l/2}$
under the rescaling. Also, the explicit factors of $R$ in the action (*)
will be rescaled by
\[
R \rightarrow ({R_s \over R})^{1/2} R
\]
Applying these transformations to the action (*), we find that the
resulting overall power of $R_s$ multiplying a term
\[
(\partial_{k_1} \cdots \partial_{k_l} h) I_n
\]
in the final action is $(R_s)^{(n+k-2)/2}$ where $2 \ge k \ge -2$ counts
the number of $+$ indices minus the number of $-$ indices on $h$.
Thus, for a given component of $h$, the terms which are finite in the
limit have $n=2-k$. It is these terms that will remain to give the
Matrix theory action, and the sum of all such terms for the component
$h_{IJ}$ must therefore equal $T^{IJ}/2$. Terms with $n<2-k$ become
infinite in the Matrix theory limit, so the sum of such terms coupling
to a given component of $h$ must be required to cancel. T$n>2-k$ vanish
in the Matrix theory limit, so knowledge of the Matrix
theory action imposes no further constraints here.
With this in mind, we may now compare the action (*) with the matrix
theory
action to find the following relations:}
\begin{eqnarray}
({1 \over 2} I_h^{00} + {1 \over 2} I_h^{ii} + {3 \over 2} I_\phi -
2I_0^0)_0
&=& 0 \nonumber\\
({1 \over 2} I_h^{00} + {1 \over 2} I_h^{ii} + {3 \over 2} I_\phi -
2I_0^0)_4
&=& 0 \nonumber\\
({1 \over 2} I_h^{00} + {1 \over 2} I_h^{ii} + {3 \over 2} I_\phi -
2I_0^0)_8
&=&T^{--} \nonumber\\
(I_h^{0i} - I_0^i)_2 &=& 0\nonumber\\
(I_h^{0i} - I_0^i)_6 &=& T^{-i}\nonumber\\
(I_h^{ij})_0 &=& 0\label{eq:constraints}\\
(I_h^{ij})_4 &=& T^{ij}\nonumber\\
({3 \over 2} I_h^{00} - {1 \over 2} I_h^{ii} - {3 \over 2} I_\phi )_0
&=& 0\nonumber\\
({3 \over 2} I_h^{00} - {1 \over 2} I_h^{ii} - {3 \over 2} I_\phi )_4
&=&
2T^{+-}\nonumber\\
( I_h^{0i} + I_0^i)_2 &=& 2T^{+i}\nonumber\\
({1 \over 8} I_h^{00} + {1 \over 8} I_h^{ii} + {3 \over 8} I_\phi + {1
\over
2} I_0^0 )_0 &=& T^{++} \nonumber
\end{eqnarray}
We will assume that the degrees at which a given current $I_x$ has
nonvanishing contributions are the same as in the $N = 1$ case. These
are the terms for which we have explicitly written constraints in
(\ref{eq:constraints}). This assumption agrees with what we know
about the nonabelian Born-Infeld action. If this assumption is
incorrect, there may be additional terms appearing in the IIA currents
at other orders which do not contribute to the matrix theory action.
Identical relations to (\ref{eq:constraints}) must hold for the
quantities coupling to higher order terms in the Taylor expansion of
the metric.
Solving the constraints (\ref{eq:constraints}), we find
\begin{eqnarray}
I_h^{00} &=& T^{++} + T^{+-} + (I_h^{00})_8 + {\cal O} (v^6)
\nonumber\\
I_h^{0i}
&=&T^{+i} + T^{-i} + {\cal O} (v^5) \nonumber\\
I_h^{ij} &=& T^{ij} +
(I_h^{ij})_8 + {\cal O} (v^6) \label{eq:result-h}\\
I_\phi &=& T^{++} - {1 \over 3} T^{+-} - {1
\over 3} T^{ii} + (I_\phi)_8 + {\cal O} (v^6) \nonumber\\
I_0^0 &=& T^{++} \nonumber\\ I_0^i &=& T^{+i} \nonumber
\end{eqnarray}
Here, we have assumed that the one-form field components
$C_0$ and $C_i$ should couple to the D0-brane charge $N/R=T^{++}$ and
spatial current ${\rm Tr}\;(\dot{X}^i)/R=T^{+i}$ so that the last two
expressions are exact. The fourth order quantities $(I_h^{00})_8$,
$(I_h^{ij})_8$, and $(I_\phi)_8$ are not completely determined by our
analysis, but they must obey the relation
\[
({1 \over 2} I_h^{00} + {1 \over 2} I_h^{ii} + {3 \over 2} I_\phi)_8 =
T^{--}
\]
From the $N = 1$
results (\ref{eq:n1}) we expect that
\begin{eqnarray*}
(I_h^{00})_8 & = & \frac{3}{2} T^{--}+ (I_h^{00})_{8c} \nonumber\\
(I_h^{ii})_8 & = & 2 T^{--}+ (I_h^{ii})_{8c} \nonumber\\
(I_\phi)_8 & = & -\frac{1}{2} T^{--}+ (I_\phi)_{8c}\nonumber
\end{eqnarray*}
where $(I)_{8c}$ are quantities of order $v^4$ which contain
commutators or fermions and which vanish in the $N = 1$ case of a
single spinless graviton considered in the previous subsection.
Additional information about the currents should follow from the
conservation of the D0 brane stress-energy tensor $I_h^{\mu
\nu}(\vec{x})$ which is defined in terms of the moments $I_h^{\mu \nu
(k_1 \cdots k_n)}$. As discussed in \cite{mvr}, the relation $D_\mu
I_h^{\mu \nu} = 0$ implies
\[
\partial_t I_h^{0 \mu (k_1 \cdots k_n)} = I_h^{k_1 \mu (k_2 \cdots k_n)}
+ \dots + I_h^{k_n \mu (k_1 \cdots k_{n-1})}
\]
In particular, $(I_h^{ij})_8$ should be precisely determined by
\[
(I_h^{ij})_8 = (\partial_t T^{+i(j)} + \partial_t T^{-i(j)})_8.
\]
So far, we have only dealt with the case of a background metric $h$,
dilaton field $\phi$ and R-R 1-form field $C$.
The same sort of analysis can be applied for backgrounds
having nonvanishing 2-form $B$ or 3-form $C^{(3)}$ fields and their
duals, and in fact the analysis is simpler in these cases.
In order to describe nontrivial background antisymmetric tensor
fields, we must generalize the relations (\ref{eq:relations}) which
connect the IIA background fields to the 11D background 3-form field
through the Seiberg limit. The $B$ and $C^{(3)}$ fields are related
to components of the 3-form field through
\begin{eqnarray*}
B_{0i} & = & \hat{A}_{10 \; 0 i} = A_{+ -i}\\
B_{ij} & = & \hat{A}_{10 \; ij} = \frac{\alpha}{\sqrt{2}} A_{+ ij} -
\frac{1}{ \alpha \sqrt{2}} A_{-ij}\\
C^{(3)}_{0ij} & = & \hat{A}_{0 \; ij} = \frac{\alpha}{\sqrt{2}} A_{+
ij} + \frac{1}{ \alpha \sqrt{2}} A_{-ij}\\
C^{(3)}_{ijk} & = & \hat{A}_{ijk} = A_{ijk}
\end{eqnarray*}
The constraints on the string and D2-brane currents analogous to
(\ref{eq:constraints}) are then
\begin{eqnarray}
(I_2^{ijk})_0 = (I_s^{0i})_0 = (3I_2^{0ij} -I_s^{ij})_2 & = & 0
\nonumber\\
(I_s^{0i})_4 & = & 3J^{+ -i}\nonumber \\
(I_2^{ijk})_4 & = & J^{ijk}\label{eq:constraints-a}\\
(3 I_2^{0ij} + I_s^{ij})_2 & = & 6 J^{+
ij} \nonumber\\
(3I_2^{0ij} -I_s^{ij})_6 & = & 3 J^{-ij} \nonumber
\end{eqnarray}
from which we can determine
\begin{eqnarray}
I_s^{0i} & = & 3J^{+ -i} +{\cal O} (v^4)\nonumber \\
I_2^{ijk} & = & J^{ijk}+{\cal O}
(v^4)\label{eq:solution-a}\\
I_2^{0ij} & = & J^{+ ij} + (I_2^{0ij})_6 +{\cal O} (v^5)
\nonumber\\
I_s^{ij} & = & 3 J^{+ ij} + (I_s^{ij})_6 +{\cal O}
(v^5)\nonumber
\end{eqnarray}
where the terms $(I)_6$ on the last two lines must satisfy the final
relation in (\ref{eq:constraints-a}). In addition, conservation
relations for $I_s$ suggest that
\[
\partial_t I_s^{0i(j)} = I_s^{ji} = - I_s^{ij}
\]
from which it follows that
\[
(I_s^{ij})_6 = -3\partial_t J^{+-i(j)} = -3 J^{-ij}, \; \; \; \; \;
(I_2^{0ij})_6 = 0
\]
There are a number of comments worth making about the identifications
(\ref{eq:solution-a}). First, note that the factor of 3 appearing in
the currents $I_s$ arises from our somewhat unconventional choice of
normalization for the couplings $A_p J^p$ between a $p$-form field and
its associated current. Often, a factor of $1/p!$ is included in the
definition of this coupling. With that redefinition of the currents,
the factors of $3$ in our relations would disappear. We have chosen
our convention for the coupling to conform with previous literature on
the subject.
Next, we briefly discuss the physical interpretation of the leading
terms in (\ref{eq:solution-a}).
The leading term $J^{+ ij}$ in $I_2^{0ij}$
is the total membrane charge of the D0-brane system.
This result is the T-dual of the statement that $\int F$ on a $p$-brane
is
the total $(p-2)$-brane charge coupling to the $(p -1)$-form R-R field
\cite{Douglas}. Although one might think that this should be the only
contribution to the D2-brane charge of the system,
additional contributions may arise from the geometry of the brane
embedding
\cite{bvs,ghm,Cheung-Yin,Minasian-Moore}.
Note that while for finite $N$ the integrated membrane charge
$J^{+ ij} = {\rm Tr}\;[X^i, X^j]$ vanishes identically (since a finite
size system can have no net membrane charge) the higher moments of
the D2-brane charge can be nonvanishing and will couple to the
derivatives of the $C^{(3)}$ field.
The leading term in $I_s^{0i}$ is the net string winding charge in
direction $i$; this is simply the T-dual of the Poynting vector giving
momentum on a dual D-brane. The leading term in the current
$I_s^{ij}$ is perhaps somewhat surprising. Although this term itself
vanishes for finite size matrices, as mentioned above, the first
moment is nonvanishing. The existence of this term indicates that
there will be a coupling in the multiple D0-brane action of the form
\begin{equation}
\partial_{[i} B_{jk]} {\rm Tr}\; (X^{[i} X^j X^{k]} + {\rm fermions}).
\end{equation}
We are rather confused as to the physical origin of this term.
Indeed, this term plays a puzzling role in several related situations.
For example, after compactification on $T^3$, the term $J^{+ ij}$
should be related by a duality transformation to the NS5-brane charge
of the IIA theory \cite{grt}, which we discuss below. It may be
possible to understand the role of this term in the theory by studying
a T-dual system such as a dual multiple D3-brane configuration. We
discuss the connection with the dual theory briefly in the last
section of this paper.
Now let us consider the currents coupling to the dual fields
$\tilde{B}$ and $\tilde{C}^{(3)}$. These currents can be derived in a
fashion precisely analogous to the above argument by considering the
fields related to the dual 6-form $\tilde{A}$ of the 11D theory. We
find
\begin{eqnarray}
I_4^{0i jkl} & = & 6 M^{+ -i jkl}+{\cal O} (v^4)\nonumber \\
I_5^{ijklmn} & = & M^{ijklmn}+{\cal O}
(v^4)\label{eq:solution-ad}\\
I_5^{0ij klm} & = & M^{+ ij klm} + (I_5^{0ij klm})_6 +{\cal O} (v^5)
\nonumber\\
I_4^{ij klm} & = & 6 M^{+ ij klm} + (I_4^{ij klm})_6
+{\cal O} (v^5)\nonumber
\end{eqnarray}
where
\[
6 (I_5^{0ij klm})_6 -
(I_4^{ij klm})_6= 6M^{-ijklm}.
\]
Just as for $I_s$, conservation relations for $I_4$ suggest that
\[
\partial_t I_4^{0ijkl(m)} = I_4^{ijklm}
\]
from which it follows that
\[
(I_4^{ijklm})_6 = 6\partial_t M^{+-ijkl(m)} = -6 M^{-ijklm}, \; \; \; \;
\; (I_5^{0ijklm})_6 = 0.
\]
The leading term in $I_4^{0ijkl}$ is the net D4-brane charge
\cite{bss,grt}. This is the dual of the instanton number on a
D$p$-brane. Unfortunately, we only know from matrix theory the
components of the 5-brane current $M^{-IJKLM}$ with one $-$ index.
Thus, we can only determine the leading term in the components
$I_4^{0ijkl}$ of the D4-brane current, and we have no information
about the leading terms in the remaining components of $I_4$
or any components
of the NS5-brane current $I_5$. The absence of any known operator for
the transverse 5-brane charge $M^{+ ijklm}$ in matrix theory is an
infamous problem. No operator of this form appears in the
supersymmetry algebra \cite{bss} or in the one-loop effective
potential \cite{Dan-Wati-2}. Nonetheless, we should expect higher
moments of this operator to appear, corresponding to local transverse
5-brane charge. It has been argued that in a T-dual 3-brane picture
the desired operator is S-dual to the charge of a D5-brane
perpendicular to the 3-brane \cite{grt} (described by the operator
$J^{+ ij}$ mentioned above), although no explicit description of this
dual operator has been given. It would be nice to have a better
understanding of these terms in the multiple D0-brane action.
Finally, we consider the currents coupling to the dual field
$\tilde{C}$. We expect the IIA 6-brane current to couple to this
field. Unlike the other branes whose currents we have considered, the
IIA 6-brane does not arise in a simple fashion from the dimensional
reduction of the membrane or the 5-brane of 11D M-theory. Rather, the
IIA 6-brane represents a nontrivial metric of Kaluza-Klein monopole
form in the 11D theory. Nonetheless, in \cite{Mark-Wati-3} we found a
matrix theory description of interactions between such metrics and
0-branes. This appeared in the form of a term in the 2-body matrix
theory potential which coupled the 0-brane stress tensor to a 10D
6-brane current $S^{\mu \nu \rho \lambda \sigma \tau \upsilon}$.
Since this current already has an essentially 10D form, it is natural
to map it directly to the 6-brane current we expect in the IIA
theory. Thus, we believe that the 6-brane current of a system of many
0-branes which couples to the background $\tilde{C}$ field will be
given by
\begin{eqnarray}
I_6^{0ijklmn} & = & S^{+ijklmn} +{\cal O} (v^5) \label{eq:solution-6}\\
I_6^{ijklmn p} & = & S^{ijklmn p}+{\cal O} (v^6)\nonumber
\end{eqnarray}
where the matrix theory form of the 6-brane current is given in the
Appendix.
\subsection{Summary of results for multiple D0-brane action in IIA}
We summarize here our results for the terms in the multiple D0-brane
action which couple linearly to the background fields of type IIA
supergravity and their derivatives. The full action including all
linear terms
is given by (\ref{eq:IIA-general})
\begin{eqnarray}
S &= & S_{{\rm flat}} +
\int dt \sum_{n=0}^{\infty} {1
\over n!} \left[
\frac{1}{2}
(\partial_{k_1}\cdots
\partial_{k_n} h^{IIA}_{\mu \nu}) \; I_h^{\mu \nu (k_1 \cdots k_n)}
+ (\partial_{k_1}\cdots \partial_{k_n} \phi)
\; I_{\phi}^{(k_1 \cdots k_n)}\right.\nonumber\\
& & \hspace{1in}+ (\partial_{k_1}\cdots
\partial_{k_n} C_{\mu }) \; I_0^{\mu (k_1 \cdots k_n)}
+
(\partial_{k_1}\cdots
\partial_{k_n} \tilde{C}_{\mu \nu \lambda \rho \sigma \tau \zeta })
\; I_6^{\mu \nu \lambda
\rho \sigma \tau \zeta (k_1 \cdots k_n)}
\nonumber\\
& & \hspace{1in}+
(\partial_{k_1}\cdots \partial_{k_n}
B_{\mu \nu}) \; I_s^{\mu \nu (k_1 \cdots k_n)}
+
(\partial_{k_1}\cdots \partial_{k_n}
\tilde{B}_{\mu \nu \lambda \rho \sigma \tau}) \; I_5^{\mu \nu \lambda
\rho \sigma \tau
(k_1 \cdots k_n)} \nonumber\\
& &\hspace{1in} \left.+
(\partial_{k_1}\cdots
\partial_{k_n} C^{(3)}_{\mu \nu \lambda }) \; I_2^{\mu \nu \lambda (k_1
\cdots
k_n)}
+
(\partial_{k_1}\cdots \partial_{k_n} \tilde{C}^{(3)}_{\mu \nu \lambda
\rho \sigma }) \; I_4^{\mu \nu \lambda \rho \sigma (k_1 \cdots
k_n)}\right] \nonumber \end{eqnarray} The multipole moments of the stress tensor
$I_h$ and currents coupling to the background dilaton and R-R 1-form
field have leading terms given by (\ref{eq:result-h}). The currents
coupling to the NS-NS antisymmetric 2-form field and the R-R 3-form
field have leading terms given by (\ref{eq:solution-a}). Of the
currents coupling to the duals of these two fields, we have only been
able to identify leading term in the the moments of the 4-brane
current component $I_4^{0ijkl}$ as described in
(\ref{eq:solution-ad}). We believe that the leading terms in the
components of the 6-brane current coupling to the dual of the 1-form
field are as given in (\ref{eq:solution-6}).
We have derived these results based on our proposal in
\cite{Mark-Wati-3} for the form of the matrix theory action in weak
background fields and Seiberg's scaling argument. Our results agree
with the known terms in the $N = 1$ Born-Infeld action, and with the
known BPS charges of the multiple D0-brane system. In the following
section we give a simple test of the results for the terms coupling to
the background metric. Further possible tests, applications, and
extensions of these results are discussed in the concluding section.
\section{Tests of the action}
In this section, we test our proposals for the general background
actions through two related calculations. First, we consider the
D0-brane action in a background describing two seperated branes in a
curved space ($h_{ij} \ne 0$). We determine the masses of off diagonal
bosonic and fermionic fields and show that these exactly match the
geodesic distance to leading order in $h$ in agreement with the
constraint suggested by Douglas. Next, we consider the analogous
background in matrix theory and compute the leading order one-loop
potential between two gravitons in a curved transverse space, showing
that curved-space supergravity predictions are reproduced.
\subsection{The geodesic length criterion}
\label{sec:masses}
One of the earliest discussions of the problem of formulating a
low-energy theory for many D0-branes moving in a curved space was given
in \cite{Douglas-curved}. In that paper, Douglas argued that one of
the most basic conditions which such an action must satisfy is that in a
background corresponding to a pair of D0-branes living at points $x$
and $y$ there should be light fields with masses equal to the geodesic
length between these points. This condition, together with additional
axiomatic assumptions, was used by Douglas, Kato and Ooguri in
\cite{dko} to give the first few terms in the D0-brane action on a
general Calabi-Yau 3-fold which preserves some supersymmetry. In this
section we show that our formulation of the linearized coupling
to a weak background in the multi-D0-brane action satisfies Douglas's
geodesic length criterion.
\subsubsection{Setup}
We wish to consider a pair of D0-branes at separated points in a weakly
curved space. We assume that the transverse metric is described by a
small perturbation about a flat background, $g_{ij} = \delta_{ij} +
h_{ij}$, while the other components of the metric and the other
background fields are trivial. Without loss of generality, we choose
coordinates so that one brane is at the origin, while the other has
transverse coordinates $r^i$. The situation is described by the
multi-D0-brane action (\ref{eq:IIA-general}) with non-zero $h_{ij}$ and
fields expanded about background matrices as
\begin{eqnarray}
X^i &= &\left(\begin{array}{cc}
r^i & 0\\
0 & 0
\end{array} \right)+
\left(\begin{array}{cc}
\zeta^i & z^i\\
\bar{z}^i & \tilde{\zeta}^i
\end{array} \right)
\label{eq:background}
\\
\Theta & = &\left(\begin{array}{cc}
0 & 0\\
0 & 0
\end{array} \right)+
\left(\begin{array}{cc}
\eta^i & \chi^i\\
\bar{\chi}^i & \tilde{\eta}^i
\end{array} \right), \nonumber
\end{eqnarray}
Here, the fields $\zeta$, $z$, $\eta$, and $\chi$ represent fluctuations
about the background.
The geodesic length condition formulated by Douglas states that the
masses of the off-diagonal fields $z$ and $\chi$, which arise from
strings stretched between the separated branes, should precisely match
the geodesic length measured by the metric $h_{ij}$ between the points 0
and $r^i$. We will now compute both the geodesic length and the
oscillator masses and show that they agree.
\subsubsection{Geodesic distance}
We begin with the geodesic length between points $0$ and $r^i$. This
geodesic length is the minimum value of the length
\begin{equation}
\int_\gamma ds = \int_0^{r^i} \sqrt{g_{ij}dx^i dx^j}
\label{eq:geodesic-length}
\end{equation}
taken over all paths $\gamma$ between the two points. Because the
geodesic path is an extremum of this functional, the variation of the
length under a small variation $\delta \gamma$ of the path is of order
$(\delta \gamma)^2$. Since we are interested in changes in the length
which are linear in the background metric, we can therefore neglect
effects from the change of the geodesic path and simply evaluate the
change in the geodesic length by integrating (\ref{eq:geodesic-length})
along the straight line which is the geodesic in the flat metric.
Thus, we take
\[
x^i(\lambda) = r^i \lambda
\]
and find
\begin{eqnarray*}
d(0,r^i) &=& \int_0^1 d\lambda \sqrt{g_{ij}(\vec{x}(\lambda))\dot{x}^i
\dot{x}^j}\\
&=& \int_0^1 d\lambda \sqrt{r^2+h_{ij}(\vec{r}\lambda)r^i r^j}\\
&=& \int_0^1 d\lambda \{r + {1 \over 2r} h_{ij}(\vec{r}\lambda)r^i r^j
+{\cal O} (h^2)\}\\
&=& r + {1 \over 2r} r^i r^j H_{ij}
\end{eqnarray*}
where
\begin{eqnarray*}
H_{ij} &=& \int_0^1 d \lambda \, h_{ij}(\lambda \vec{r})\\
&=& \sum_{n=0}^\infty{1 \over (n+1)!} (r \cdot \partial)^n h_{ij}(0)
\end{eqnarray*}
This gives us the geodesic length between the two points to
linear order in the background metric. In the following section, it will
be most convenient to compare squared oscillator masses with the squared
geodesic length, given by
\begin{equation}
\label{eq:length}
d^2 = r^2 + r^i r^j H_{ij} +{\cal O} (h^2)
\end{equation}
\subsubsection{Oscillator masses}
\label{sec:oscillator-masses}
We now calculate the masses of the off-diagonal fields. From
(\ref{eq:IIA-general}), we find that the $N=2$ D0-brane action in
the case of a transverse background metric has leading terms \begin{eqnarray} S
&=& {1 \over 2R} \int dt \; {\rm Tr}( D_t X^i D_t X^i + {1 \over 2}
[X^i, X^j] [X^i, X^j] + i \Theta D_t \Theta - \Theta[\slash{X},
\Theta]) \nonumber\\ & & + {1 \over 2} \int dt \; \sum_{n=0}^\infty {1
\over n!} \partial_{k_1} \cdots \partial_{k_n} h_{ij} T^{ij(k_1 \cdots
k_n)}
\label{eq:Trans-action}
\end{eqnarray}
Here,
\begin{eqnarray}
T^{ij(k_1 \cdots k_n)} &=& {1 \over R} {\rm STr}\;
\left( \left\{ D_t X^i D_t X^j -
[X^i,
X^k] [X^k, X^j] \right.\right. \label{eq:tij}\\
& &\left.\left.\hspace{1in}
- {1 \over 4} \Theta \gamma^i [X^j, \Theta] - {1 \over
4} \Theta \gamma^j [X^i, \Theta] \right\} X^{k_1} \cdots X^{k_n}\right)
\nonumber\\ &
&
\hspace{0.3in} + \tilde{T}^{ij(k_1 \cdots k_n)}\nonumber
\end{eqnarray}
We note here that this is precisely the action for Matrix theory in a
background $h_{ij}$ to leading order in the metric. In the next section
we will use exactly this action to calculate the one-loop potential
between two gravitons, taking the same background (\ref{eq:background}),
though allowing
$\vec{r}$ to be a function of time. Such a calculation is simplest using
a gauge fixed version of the action in which we choose the background
field gauge, adding a term
\[
S_{fix} = {1 \over R} \int (-D_t X^0 + i[B^i, X^i])^2
\]
plus the appropriate ghost terms. For later convenience, we will analyze
this gauge fixed version, keeping in mind both the Matrix theory
interpretation and D0-brane interpretations.
Unlike the Matrix theory action, the complete D0-brane action contains
additional terms both in the background intependent part and coupled to
$h_{ij}$. However these cannot contribute to the quadratic action since
they contain more than two matrices (eg $\dot{X}^i$, $F_{ij}$, $\Theta$)
in which there are no entries depending only on background fields.
Similarly, the terms $\tilde{T}^{ij(k_1 \cdots k_n)}$, whose form has
not been worked out for $n>1$ involve at least two fermions $\Theta$ and
one power of $\dot{X}$ or $[X^i, X^j]$ and so contribute only cubic and
higher order terms to the action, irrelevant for determining the
oscillator masses or computing the one loop potential.
We now replace $X$ and $\Theta$ in the action with the matrices given
in (\ref{eq:background}), and write down the terms in the action
quadratic in the off
diagonal fields $z$ and $\chi$. It turns out that the symmetrization
prescription for ordering the matrices in $T^{ij(k_1 \cdots k_n)}$ is
very important here, since most of the orderings give no contribution to
the quadratic terms we are interested in. For example, the first term
in (\ref{eq:tij}) contains a term
\begin{eqnarray*}
\lefteqn{{1 \over 2 n!} (\partial_{k_1} \cdots \partial_{k_n} h_{ij}) \;
\;
{\rm STr}\;(\dot{X^i} \dot{X^j} \; X^{k_1} \cdots X^{k_n})} \\
&=& {1 \over 2 n!} (\partial_{k_1} \cdots \partial_{k_n} h_{ij})\; \;
{1 \over
n+1}
\sum_{m=0}^n {\rm Tr}\;(\dot{X}^i X^{k_1} \cdots X^{k_m} \dot{X}^j
X^{k_{m+1}}
\cdots X^{k_n})
\end{eqnarray*}
for which only the $m=0$ and $m=n$ terms contribute to the quadratic
action in the off diagonal field $z$. Summing over $n$, this
contribution gives
\begin{eqnarray*}
& &\sum_{n=0}^\infty {1 \over (n+1)!} (\partial_{k_1} \cdots
\partial_{k_n}
h_{ij}) (\dot{\bar{z}}^i \dot{z}^j ) r^{k_1} \cdots r^{k_n}\\
&=&\dot{\bar{z}}^i \dot{z}^j \sum_{n=0}^\infty {1 \over (n+1)!} (r
\cdot
\partial)^n h_{ij}
\\
&\equiv&H_{ij} \dot{\bar{z}}^i \dot{z}^j
\end{eqnarray*}
Note that the quantity $H_{ij}$ is simply the function $h_{ij}$
integrated over the straight line trajectory between $0$ and $\vec{r}$,
\[
H_{ij} = \int_0^1 h_{ij}(\lambda \vec{r}) d \lambda
\]
which also appeared at first order in the geodesic distance formula
(\ref{eq:length}). Using this definition, it is straightforward to write
down the remaining terms in the quadratic actions for each of the off
diagonal fields.
\vspace{0.15in}
\noindent {\bf Bosonic Terms}
\vspace{0.15in}
In exactly the same way as for the terms just derived, we find that the
complete set of terms to leading order in $h$ for the nine transverse
bosonic
fields is
\[
S_B = -\bar{z}^i\{(\partial_t^2 + r^2) (\delta_{ij} + H_{ij}) + r^k r^l
H_{kl} \delta_{ij} - r^i r^k H_{kj} - H_{ik} r^k r^j \}z^j
\]
Note that the matrix $(\delta_{ij} + H_{ij})$ multiplies all terms not
containing $h$. Since the remaining terms are already of order $h$, if
we
redefine $z^i$ to eliminate this factor in the first terms, the
remaining
terms will only be changed at second order in $h$. After such a field
redefinition, the kinetic term is proportional to the identity, and the
squared oscillator masses are therefore given by the eigenvalues of the
constant matrix
\begin{eqnarray*}
{\bf M} &=& r^2 \left\{ (1+H_{rr}){\rlap{1} \hskip 1.6pt \hbox{1}}_{9 \times 9}
+ \left( \begin{array}{cccc} 0 & \cdots & 0 & H_{1r}\\ \vdots & & &
\vdots \\
0 & & 0 & H_{8r} \\ H_{1r} & \cdots & H_{8r} & -2H_{rr} \end{array}
\right) +
{\cal O} (h^2) \right\}
\end{eqnarray*}
Here, to simplify the formulae, we have made a rotation so that
$\vec{r}$ lies in the $9$ direction, which we refer to using the index
$r$. It is straightforward to solve directly for the eigenvalues and
eigenvectors of this mass matrix, and one finds to this order that the
masses are
\[
m_1^2 = \cdots = m_7^2 = r^2(1 + H_{rr}), \; \; m_8^2 = r^2(1 +
\sqrt{H_{rr}^2
+
H_{ri} H_{ir}}), \; \; m_9^2 = r^2(1 - \sqrt{H_{rr}^2 + H_{ri} H_{ir}})
\]
where the index $i$ is summed from 1 to 8. The oscillators with masses
$m_9$ and $m_8$ correspond to directions lying in the plane defined by
the $r$ direction and the perpendicular vector $H_{ir}$. The remaining
oscillators correspond to the directions perpendicular to these and have
masses which precisely match the geodesic distance (\ref{eq:length}) to
leading order
in $h$.
The agreement between the masses of these perpendicular oscillators
and the geodesic distance is precisely the criterion used by Douglas
et al.
in \cite{Douglas-curved,dko} to constrain the leading terms in
multiple D0-brane action on certain classes of manifolds. The fact
that the oscillators corresponding to fields not perpendicular to the
separation have different masses is also expected. In the non-gauge
fixed theory, the off-diagonal fields in the direction of the
separation between the branes simply give a combination of a gauge
rotation of the system and a relative motion of the D0-branes along a
flat direction. This effect explains the failure of the masses $m_8,
m_9$ to satisfy Douglas's criterion.
\vspace{0.15in}
\noindent {\bf Gauge Field}
\vspace{0.15in}
For a time independent $\vec{r}$, there is no mixing between gauge
field and the other bosonic oscillators in the quadratic action, and we
find that the quadratic terms involving the off diagonal field $z_0$ are
simply
\[
S_A =-\bar{z}_0 \{ \partial_t^2 + r^2 + r^2H_{rr} \} z_0
\]
Hence, the off diagonal gauge field also has mass equal to the geodesic
distance to leading order in $h$,
\[
m_0^2 = r^2 + H_{rr} \,.
\]
\vspace{0.15in}
\noindent {\bf Ghost Fields}
\vspace{0.15in}
Since our gauge fixing term does not depend on the background
metric, the off diagonal ghost fields will have a mass given by $m_g^2 =
r^2$, as in the flat space case, with action
\[
S_G = -\bar{c} \{ \partial_t^2 + r^2 \} c
\]
\vspace{0.15in}
\noindent {\bf Fermionic Fields}
\vspace{0.15in}
Proceeding in the same way for the quadratic fermion action, we find
that the action quadratic in the off diagonal field $\chi$ is
\[
S_F =-\bar{\chi}_\alpha \{i\partial_t + \gamma^i_{\alpha \beta} (r^i +
{1 \over
2} H_{ij} r^j) \} \chi_\beta
\]
Thus, in the presence of a background metric, the quadratic fermion
action is only changed by a shift
\[
r^i \rightarrow r^i + {1 \over 2} H_{ij} r^j
\]
so the sixteen fermion fields $\chi$ have a mass squared matrix given
by
\[
M_f^2 = {\rlap{1} \hskip 1.6pt \hbox{1}} (r^2 + H_{ij} r^ir^j + {\cal O} (h^2)) = {\rlap{1} \hskip 1.6pt \hbox{1}}
r^2(1 +
H_{rr} +
{\cal O} (h^2))
\]
We see that all the fermionic oscillators have a mass which reproduces
the geodesic distance to leading order in $h$.
\vspace{0.15in}
\noindent {\bf Summary}
\vspace{0.15in}
To summarize, we have found eight complex bosons (including the gauge
field) and sixteen real fermions with masses equal to the geodesic
distance between $0$ and $\vec{r}$. Additionally, there are complex
bosons with $m^2 = r^2(1 + \sqrt{H_{rr}^2 + H_{ri} H_{ir}})$ and $m^2 =
r^2(1 - \sqrt{H_{rr}^2 + H_{ri} H_{ir}})$ and two complex ghosts with
$m^2 = r^2$.
Thus, the sum of the squared masses weighted by the number of
degrees of freedom is identical for the fermions and the bosons
(including the ghosts with negative weight), and for both sets of
fields, the average mass squared per degree of freedom is exactly the
geodesic distance (\ref{eq:length}). We have now seen that the geodesic
distance
criterion is precisely realized in the proposed multi-D0-brane action.
\subsection{Graviton interactions}
In this section, we use the general background Matrix theory action to
study the interactions between two gravitons with unit momentum around a
lightlike circle in a weakly curved space. In various cases, we will
compute the one-loop effective action to first order in the metric and
compare with the interactions expected from DLCQ supergravity with a
curved background.
The relevant matrix theory action and background are exactly the same as
those considered for D0-branes in the previous section. In this case,
however, we do not wish to restrict to gravitons which are fixed in the
transverse space, so we allow $\vec{r}$ to be a function of time.
As for the case of flat space Matrix theory, we should require that
our background matrices satisfy the equations of motion. For block
diagonal backgrounds, the equations of motion decouple for each block,
so for our case, we require that $\vec{r}(t)$ satisfy the equations of
motion derived from the $U(1)$ action (\ref{graviton}). For a metric
which is non-trivial only in the transverse directions, the equations of
motion are
\begin{equation}
\label{eq:eom}
\ddot{r}^i = \dot{r}^k \dot{r}^l g^{im} (\vec{r}) \{ {1 \over 2}
\partial_m
g_{kl} (\vec{r}) - \partial_k g_{ml} (\vec{r}) \}
\end{equation}
This is just the equation for a free non-relativistic particle moving in
a curved space. We will consider two simple cases of trajectories which
trivially satisfy these equations of motion. First, we may consider the
static case $\vec{r}(t) = \vec{r}(0)$. The second case is one for which
the metric has a flat direction $i$ (so that $h_{ij} =
\partial_ig_{jk} = 0$)
and we take the particle to have some velocity in this direction. In
this case, the right hand side of (\ref{eq:eom}) vanishes, so that
$r^i(t) = r^i + v^it$.
\subsubsection{Supergravity predictions}
Before proceeding with the matrix theory calculation, we would like to
see what supergravity predicts for the interaction potential between two
gravitons in a weakly curved space.
As above, we start with a static metric $g_{ij}(\vec{x}) = \delta_{ij} +
h_{ij}(\vec{x})$ which is assumed to satisfy the source-free Einstein
equations
\begin{equation}
\label{eq:Ricci}
0 = R_{ij} ={1 \over 2}(\Delta h_{ij} - \partial_i \partial_k h_{kj} -
\partial_j \partial_k h_{ki} + \partial_i \partial_j h_{kk}) + {\cal
O}(h^2)
\end{equation}
We may find the potential between a pair of gravitons in this space by
treating one as a source for a perturbation about the metric $g$ and
reading off the potential from the probe action (\ref{graviton}),
keeping only terms arising from the perturbation in the original metric
due to the presence of the source graviton.
For our source, we choose the graviton which sits at the origin of the
transverse space with unit momentum in the compact direction. The
stress-energy tensor for this particle still has only a single
non-vanishing component,
\[
T^{++} = T_{--} = {1 \over 2 \pi R^2} \delta(\vec{x}),
\]
The presence of this graviton will result in a perturbation of the
metric $g_{ij}$ which we denote by $\gamma_{ij}$. The fact that $T_{--}$
is the only non-vanishing component of the stress-energy tensor
simplifies things considerably, and as with the flat space case, we may
solve the Einstein equation taking only the component $\gamma_{--}$ to
be non-zero. In this case, the condition that the perturbed metric $g +
\gamma$ should continue to satisfy the Einstein equation with source $T$
reduces at leading order in $\gamma$ to the covariant Laplace equation,
\begin{equation}
g^{ij}\nabla_i \nabla_j \gamma_{--} = g^{ij} (\partial_i \partial_j -
\Gamma^k_{ij} \partial_k ) \gamma_{--} = 2\kappa_{11}^2 T_{--}
\label{eq:laplace}
\end{equation}
(see, for example \cite{Weinberg}). We are only interested in the
solution at leading order in $h$ (the original background metric), so we
expand
\[
\gamma_{--} \equiv \gamma_0 + \gamma_1 + {\cal O}(h^2)
\]
Here $\gamma_0$ is the part independent of $h$, equal to the flat space
solution (ignoring non-numerical constants)
\begin{equation}
\label{eq:flat}
\gamma_0 = {15 \over 2 r^7}
\end{equation}
while $\gamma_1$ is the part linear in $h$.
In the case where $g_{ij}$ is the metric corresponding to some choice of
coordinates on flat space, the exact solution to (\ref{eq:laplace}) must
be given by a covariant version of (\ref{eq:flat}), replacing $r$ with
the geodesic distance $d$ between $0$ and $r$. In this case, using
(\ref{eq:length}), we have
\begin{equation}
\label{eq:hlinear}
\gamma = {15 \over 2} \{ {1 \over r^7} - {7 r^i r^j \over 2r^9} \int_0^1
d \lambda h_{ij}(\lambda \vec{r})\} + {\cal O}(h^2)
\end{equation}
In fact, this solves (\ref{eq:laplace}) to leading order in $h$ in any
case where the metric $g$ is Ricci-flat as may be verified explicitly by
substitution. The $h$ independent part of (\ref{eq:laplace}) reads
\[
\partial^2 \gamma_0 = 2\kappa_{11}^2 T_{--},
\]
and is just the statement that $\gamma_0$ is the solution for flat
space. The linear terms in $h$ in equation (\ref{eq:laplace}) read
\[
\partial^2\gamma_1 = h_{ij}\partial_i \partial_j \gamma_0 + ( \partial_i
h_{ik} - {1 \over 2} \partial_k h_{ii}) \partial_k \gamma_0
\]
and substituting for $\gamma_0$ and $\gamma_1$ from (\ref{eq:hlinear})
it is not hard to check that this holds, making use of the
Ricci-flatness condition (\ref{eq:Ricci}), the identity
\[
\partial_\lambda h(\lambda \vec{r}) = (r \cdot \partial) h(\lambda
\vec{r}),
\]
and various integrations by parts. This is done most easily by choosing
coordinates so that $h$ satisfies the harmonic gauge condition in which
\[
\partial_i h_{ij} = { 1 \over 2} \partial_j h_{ii}, \;\;\;\;R_{ij} =
\partial^2 h_{ij} + {\cal O}(h^2)
\].
Thus, to leading order in the backgrounds we are considering, the metric
perturbation due to the presence of a graviton at the origin is simply
\[
\gamma_{--}(\vec{x}) = {15 \over 2 d^7(\vec{x})}
\]
where $d$ is the geodesic distance between 0 and $\vec{x}$. Recalling
the action (\ref{graviton}) for the probe graviton moving in the metric
produced by this source, we see that to leading order in the background,
the curved space graviton-graviton potential is simply
\[
V = -{15 \over 16} {v^4 \over d^7}
\]
In the next sections, we will carry out the graviton potential
calculation in Matrix theory to compare with this supergravity
prediction.
\subsubsection{Static case}
First, we consider the static case in which both gravitons have zero
transverse velocity. In this case, as in flat space, we expect the
potential to vanish at leading order in the transverse metric, as shown
above.
For this case, we may directly apply the results of section
\ref{sec:masses}. The complete action quadratic in the off diagonal
fields is
\begin{eqnarray}
S&=&S_B+S_A+S_F+S_G \nonumber \\
S_B &=& -\bar{z}^i\{(\partial_t^2 + r^2) (\delta_{ij} + H_{ij})
\nonumber \\
& &\hspace{0.5in} + (r^k r^l
H_{kl})\delta_{ij} - r^i r^k H_{kj} - H_{ik} r^k r^j \}z^j \nonumber \\
S_A &=& -\bar{z}_0 \{ \partial_t^2 + r^2 + r^i r^j H_{ij} \} z_0
\label{eq:stat-action}\\
S_F &=& -\bar{\chi}_\alpha \{i\partial_t + \gamma^i_{\alpha \beta} (r^i
+ {1 \over 2} H_{ij} r^j \} \chi_\beta \nonumber \\
S_G &=& -\bar{c} \{ \partial_t^2 + r^2 \} c \nonumber
\end{eqnarray}
As argued above, this leads to eight complex bosons (including the gauge
field) and sixteen real fermions with masses equal to the geodesic
distance between $0$ and $\vec{r}$ as well as complex bosons with $m^2 =
r^2(1 + \sqrt{H_{rr}^2 + H_{ri} H_{ir}})$ and $m^2 = r^2(1 -
\sqrt{H_{rr}^2 + H_{ri} H_{ir}})$ and two complex ghosts with $m^2 =
r^2$.
The vanishing of the one loop potential to leading order in $h$ is
ensured by the fact that the sum of the squared masses weighted by
number of degrees of freedom is identical for the fermions and bosons
(including ghosts weighted by -1). This follows since the one loop
effective action depends only on the oscillator masses and is given by:
\begin{eqnarray*}
e^{i\Gamma_{1 loop}}&=&\det{}^{-8}(\partial_t^2 + r^2(1 + H_{rr}))
\det{}^{8}(\partial_t^2 + r^2(1 +
H_{rr})) \det{}^{2}(\partial_t^2 + r^2)\\
& &\det{}^{-1}(\partial_t^2 + r^2(1 + \sqrt{H_{rr}^2 + H_{ri}
H_{ir}}))\det{}^{-1}(\partial_t^2 + r^2(1-\sqrt{H_{rr}^2 + H_{ri}
H_{ir}}))\\
&=&1 + {\cal O} (h^2)
\end{eqnarray*}
Hence, in the static case, we find agreement with our expectations
from supergravity.
\subsubsection{Velocity dependent potential}
We now consider the case of two gravitons with relative velocity.
Here, we assume that the particle with initial position $\vec{r}$ moves
in a direction $\vec{v}$ which is perpendicular to $\vec{r}$ and in
which the metric is flat. In this case, we have shown above that
supergravity predicts a potential
\[
-{15v^4 \over 16 d^7}
\]
where $d$ is the geodesic separation distance.
To simplify our calculations, we rotate coordinates so that $\vec{r}$
and $\vec{v}$ lie in coordinate directions which we denote by indices
$r$ and $v$. Thus
\[
g_{vv}(\vec{x}) = 1, \; g_{vr}(\vec{x})= g_{vi}(\vec{x}) =
\partial_vg_{ij}(\vec{x}) = 0
\]
These equations ensure that the matrix theory equations of motion
(\ref{eq:eom})
are satisfied for the trajectory $\vec{r}(t) = \vec{r} + \vec{v}t$ that
we are considering. Expanding the action (\ref{eq:stat-action}) about
this background, we
find that the quadratic action for the off diagonal fields is equal to
the action for the static case (where $r^i$ is interpreted as
$r^i(t)=r^i + v^it$) plus extra terms:
\[
S_v = 2i\bar{z}^iv^iz^0 - 2i\bar{z}^0v^iz^i
\]
Note that these terms (in which the background appears as
$\dot{\vec{r}}$ rather than just $\vec{r}$) come only from the flat
space action, since in the metric dependent terms, $\dot{X}^i$ only
appears coupled to $h_{ij}$ while $\dot{r}^i h_{ij} = 0$. As a result,
the remaining calculation is almost identical to the flat space
calculation, performed in \cite{DKPS}.
To see this,
we note that the
complete action in this case may be written (eliminating a factor
$(\delta_{ij} + H_{ij})$ as above to diagonalize the boson kinetic term)
\begin{eqnarray*}
S&=&S_B+S_A+S_v+S_F+S_G\\
S_B &=& -\bar{z}^i\{(\partial_t^2 + d^2 + v^2t^2) \} z^i\\
& &\hspace{0.5in} + \bar{z}^r\{r^2H_{ri}\}z^i +
\bar{z}^i\{r^2H_{ir}\}z^r + \bar{z}^v\{ H_{ri}rvt\}z^i +
\bar{z}^i\{H_{ir}rvt\}z^v \\
S_A &=& -\bar{z}_0 \{ \partial_t^2 + d^2 + v^2t^2 \} z_0\\
S_v &=& 2i\bar{z}^iv^iz^0 - 2i\bar{z}^0v^iz^i\\
S_F &=& -\bar{\chi}_\alpha \{i\partial_t + \gamma^i_{\alpha \beta} (d^i
+ v^it ) \} \chi_\beta \\
S_G &=& -\bar{c} \{ \partial_t^2 + d^2 \} c\\
& &\hspace{0.5in} + \bar{c} \{H_{rr}\} c
\end{eqnarray*}
where we have defined
\[
d^i = r^i + {1 \over 2} H_{ij} r^j
\]
so that $d^2$ is the squared geodesic distance. Apart from the terms
with explicit factors of $H$ in the boson and ghost actions, this is
exactly the flat space action with $r^i$ replaced by $d^i$. Recalling
the flat space calculation, we see that at zeroth order in $H$, the
oscillators $z^i$, $i=1,\cdots, 7$, and $z^r$ are degenerate with mass
$d^2+v^2t^2$ while $z^v$ and $z^0$ combine into oscillators with
non-degenerate masses $(d^2 + v^2t^2 \pm 2v)$. Adding the perturbation
\begin{equation}
\label{pert}
\bar{z}^r\{r^2H_{ri}\}z^i + \bar{z}^i\{r^2H_{ir}\}z^r + \bar{z}^v\{
H_{ri}rvt\}z^i + \bar{z}^i\{H_{ir}rvt\}z^v
\end{equation}
we note that the last two terms do not affect the masses at leading
order in $h$, since if we change coordinates to diagonalize the
leading order mass matrix, these contribute only to non-diagonal
matrix elements connecting eigenvectors of different mass. (Recall
from basic perturbation theory that for eigenvectors $|A_1\rangle,
\cdots |A_n\rangle$ with degenerate zeroth order eigenvalues and
eigenvectors $|B_1\rangle, \cdots |B_m\rangle$ with non-degenerate
zeroth order eigenvalues that the first order shift in the eigenvalues
for $|B_i\rangle$ come only from the matrix element $\langle
B_i|M|B_i\rangle$, while the first order eigenvalues for the space
spanned
by $|A_i\rangle$ are determined only by the submatrix $\langle
A_i|M|A_j\rangle$).
The first two terms in (\ref{pert}) have the same effect as they did
in the static case, to shift two of the degenerate boson masses by
\begin{eqnarray*}
\Delta m^2 &=& - H_{rr} \pm \sqrt{H_{rr}^2 + H_{ri}H_{ir}}\\
&\equiv&-H_{rr} \pm G
\end{eqnarray*}
Meanwhile, the perturbation in the ghost action shifts the two ghost
masses by
\[
\Delta m_g^2 = - H_{rr}
\]
Defining
\[
F(x) = det(\partial_t^2 + (d^2 - x) +v^2t^2)
\]
we find that since all of the mass shifts are time independent, the one
loop effective action is simply
\begin{eqnarray*}
e^{i\Gamma} &=& e^{i \Gamma_d} {F(H_{rr})F(H_{rr}) \over F(H_{rr} +
G)F(H_{rr} - G)}\\
&=& e^{i \Gamma_d}(1 + {\cal O} (h^2))
\end{eqnarray*}
where $\Gamma_d$ is the flat-space potential with $r$ replaced by the
geodesic length $d$. We conclude that the leading order one loop
potential is simply given by
\[
V=-{15 \over 16} {v^4 \over d^7}
\]
as predicted by supergravity.
\section{Discussion}
In this paper we have derived the leading terms in the multiple
D0-brane action in a weakly curved background metric, dilaton, and
antisymmetric tensor background fields. This action was derived by
using Seiberg's scaling arguments on an action we recently proposed
for the M(atrix) model of M-theory in weak background fields. We
found explicit forms for the IIA stress-energy tensor of a multiple
D0-brane system, as well as the components of D2-brane, D4-brane,
D6-brane and fundamental string currents which couple to the
background R-R and B fields of the IIA theory. We tested our action
by verifying that it satisfies Douglas's geodesic length criterion.
We also showed that the corrections to the one-loop effective
potential between a pair of individual 0-branes correctly reproduce
curved space supergravity results.
The results presented in this paper give for the first time a
systematic description of the linear coupling between a system of
multiple D-branes and background supergravity fields. There are a
number of previous discussions of this sort of action in the
literature to which our results can be related.
In \cite{dko}, Douglas, Kato and Ooguri used
Douglas's geodesic length criterion
and other axioms including an
assumption of supersymmetry
to constrain the
form of the multiple D3-brane action on a transverse K\"ahler
manifold. They compute the first few correction terms in terms of the
curvature tensor $R_{i \bar{j}k \bar{l}}$ of the background.
They show that the first term, corresponding to our coupling
\begin{equation}
(\partial_{k_1 k_2} h_{ij}) I_h^{ij(k_1 k_2)}
\end{equation}
is uniquely determined by the geodesic length criterion. The linear
part of the term they find indeed has the same form $\partial^2 h {\rm
Tr}\; X^2 (F^2 + \dot{X}^2)$ as our result for this term. Their
approach is not able to uniquely determine the higher order
terms, but our results are compatible with the general structure of
the linear parts of the structure they find at higher order. It is
interesting that they are able to determine the form of some of the
quadratic couplings in their work as well as the linear terms.
It is natural to try to extend the results of this paper to determine
the quadratic and higher order couplings between a system of many
D-branes and the supergravity background fields. In
\cite{Mark-Wati-3} it was suggested that the coupling to $n$th order
terms in the background could be determined by a $n$-loop calculation
in matrix theory. The results of Okawa and Yoneya on 3-graviton
scattering in matrix theory \cite{Okawa-Yoneya,Okawa-Yoneya-2} seem to
indicate that this may work at least to quadratic order in general
backgrounds, although there are indications
\cite{deg2,Sethi-Stern-2,Lowe-constraints} that there may be problems
with extending this approach to higher order. In any case, even the
two-loop calculation would be quite challenging to work out for
completely general background configurations, so it would be nice to
find a simpler approach. The terms coupling the open string fields of
the D-brane system to any number of bulk supergravity fields can of
course in principle be determined by a perturbative string
calculation. These calculations are quite complicated, however, even
for the linear coupling terms discussed in this paper. (For recent
work computing the curvature squared terms in the single D-brane
action see \cite{bbg}). Furthermore,
the results we have given here extend to arbitrary derivatives in the
background fields and contain a correspondingly arbitrary number of
D-brane fields. It is difficult to imagine reproducing such results
from perturbative string theory. Despite these difficulties inherent
in a systematic derivation of the higher order coupling terms, it may
be that the symmetries of the theory and the geodesic length criterion
are sufficient to determine the structure of some of the higher-order
terms. The results of \cite{dko} on K\"ahler manifolds indicate that
it is indeed possible to learn something about about the higher order
terms using this approach. In this paper we have found that the
combinatorial structure of the higher moment terms is crucial in
fixing the masses of off-diagonal strings in accord with the geodesic
length criterion. This indicates that this condition will place
strong constraints on the possible form of the higher order couplings
to the background. Another approach which might help extend the
results here to higher order is to find the symmetry principle which
corresponds to general coordinate invariance for the multiple D0-brane
system. Because the coordinates enter the theory as matrices, there
are ordering ambiguities in determining how the operators describing
the D0-brane system transform. If this symmetry could be understood
in a systematic fashion, it would quite possibly uniquely determine
the higher order couplings of the multiple D-brane system to general
backgrounds.
The approach we have taken here to describing matrix theory and
multiple D0-brane systems in curved backgrounds is to assume that the
action in a weak background can be written as a matrix quantum
mechanics theory with a systematic expansion in powers of the
background field strength. This approach certainly seems valid for
linear couplings to the background. It is not clear, however, that
such an approach can be extended to all orders. In
\cite{dos,Douglas-Ooguri}, in fact, it was argued that even on simple
manifolds like K3 or ALE spaces it may not be possible to describe
DLCQ M-theory using a finite size matrix quantum mechanics theory. We
are not sure at this point how the approach we have taken here fits
with the results of those papers. One possibility is that the
perturbation expansion we are considering in powers of the background
field will not converge to a well-defined theory when higher order
terms are taken into account and the background is of the form
considered in \cite{dos,Douglas-Ooguri}. Indeed, the results of those
papers indicate that the expansion in weak backgrounds may break down
at quadratic order in the curvature of the background\footnote{Thanks
to H.\ Ooguri for correspondence on this point.}. Another possibility is that
either the restriction to light-front coordinates in M-theory or
the fact that in the Seiberg limit the scale of the metric structures
of interest becomes smaller than the string length may have a subtle
effect on the relationship between the supergravity and open string
descriptions of graviton interactions, leading to some modification in
the conclusions of \cite{dos,Douglas-Ooguri}. It is clearly a very
important question whether a good low-energy description of D0-branes
can be given which maps to M-theory in the Seiberg limit, but we leave
a further resolution of this issue to further work.
In this paper we have focused on the action for a system of D0-branes
in a weak supergravity background. It is possible, however, to
T-dualize the action we have given here to get an action for
D$p$-branes of arbitrary dimension in a weak background. One
particularly interesting case is that of D3-branes. The T-duals of
the currents $I_x$ we have determined in this paper are linear
combinations of the operators in the ${\cal N} = 4,$ $D = 3 + 1$ super
Yang-Mills theory on the world volume of a system of many D3-branes
which lie in the short representations of the $SU (2, 2 | 4)$
superconformal symmetry group of the theory. These operators, which
couple linearly to the supergravity background fields, play a crucial
role in the simplest version of Maldacena's AdS/CFT correspondence
\cite{Maldacena-conjecture}. For the fields associated with the
lowest partial waves of bulk fields in the AdS space, some of the
corresponding operators were found from the Born-Infeld action in
\cite{Klebanov-absorption,gkt,flz,Das-Trivedi}. All the other lowest
partial wave operators are in principle determined by supersymmetry
and group theory from the weight 2 operator $ {\rm Tr}\; X^{(i}
X^{j)}$, following \cite{krn,Gunaydin-Marcus,gkp-2,Witten-AdS1} and
related work. Although these operators have played a fundamental role
in understanding the detailed structure of the AdS/CFT correspondence,
only a few of these operators have been described explicitly in terms
of the component fields in the D3-brane world-volume theory. In
addition, to date no general method for understanding the structure of
the higher partial wave operators (which correspond to the higher
moments of our currents $I_x^{\cdots (\cdots)}$ through T-duality and
which are related through supersymmetry to the higher weight chiral
primary operators ${\rm Tr}\; X^{\{i_1} \cdots X^{i_k\}}$) has
been presented in the literature, although some discussion of
particular operators of this type was given in, for example,
\cite{gkt,dfs2}. In a separate paper we will discuss more details of
the connection between the results presented here and the operators
which are used in the AdS/CFT correspondence for 3-branes.
A feature which emerges from the results of
\cite{Mark-Wati,Dan-Wati-2,Mark-Wati-3} and the extension of this
work in the present paper is the characteristic form of the higher
moments of the currents $I_x$ which couple to the derivatives of the
background supergravity fields. In general, we find that the $n$th
moment of the current $I_x$ has contributions of the form
\begin{equation}
I_x^{(i_1 \cdots i_n)} = {\rm Sym} \; (I_x; X^{i_1}, X^{i_2}, \ldots
X^{i_n}) + I_{x ({\rm fermion})}^{(i_1 \cdots i_n)}
\end{equation}
where the first term on the RHS is a bosonic contribution to the
higher moment given by a symmetrized trace of the polynomial giving
the monopole moment of $I_x$ with a product of $n$ $X$'s, and the
second term contains fermionic contributions to the higher moment.
The fact that all the monopole moments of the currents as well as
their higher moments can be expressed in a symmetrized trace form is
reminiscent of Tseytlin's suggestion \cite{Tseytlin} for using the
symmetrized trace to resolve ordering ambiguities in the nonabelian
Born-Infeld action. Indeed, the components $T^{--}$ of the 11D stress
tensor are precisely the symmetrized $F^4$ terms appearing at fourth
order in the nonabelian Born-Infeld action proposed by Tseytlin. This
structure may be helpful in trying to predict the form of higher-order
terms in the action without doing explicit matrix theory or string
theory calculations. This structure is also helpful in understanding
previous work in which higher partial waves of operators on the
D3-brane play a role. In particular, in \cite{gkt} the rate of
absorption of higher partial waves of minimally coupled scalars in an
extremal D3-brane background was computed in supergravity and compared
to a D3-brane world-volume calculation. While the analogous
calculations for $s$-wave absorption can be matched precisely
including numerical coefficients, for partial waves with $l > 1$ a
discrepancy was found in \cite{gkt} between the results of these two
calculations. The authors suggested that this discrepancy might arise
because the higher partial wave operators were not correctly
normalized. The results we have given here suggest by T-duality that
they indeed used the correct normalization, but that the operators
they used should have a symmetrized trace with respect to orderings of
the fields. This additional information seems to help resolve the
discrepancy found in \cite{gkt}\footnote{Thanks to Steve Gubser and Igor
Klebanov for discussions on this point}.
\junk{For example, the operators of the
form $ O ={\rm Tr}\; F F X X$ appearing as part of the second partial
wave operator ${\rm Tr}\; F^2 (X^i X^j -1/6 \delta^{ij} X^2)$ should be
replaced by the symmetrized operator
\begin{equation}
O ={\rm STr}\; FF X X =
\frac{2}{3} {\rm Tr}\; FF X X +
\frac{1}{3} {\rm Tr}\; FXF X .
\end{equation}
Since the outgoing particles from the vertices associated with the two
terms on the RHS of this expression can be distinguished at large $N$,
the single term $O$ carrying a factor of 1 should be replaced by two
vertices carrying factors of $(2/3)^2 = 4/9$ and $(1/3)^2 = 1/9$
respectively. Together, these two vertices seem to cancel the factor
of 9/5 discrepancy found in \cite{gkt} for $l = 2$, although there are
several recalcitrant factors of 2 appearing in the analysis which
complicate this story. A similar argument gives
combinatorial factors which seem to fix the discrepancy in the case $l =
3$,
again up to some rogue factors of 2.}
A more detailed
discussion of the resolution of this problem will be described in a
future publication.
\section*{Acknowledgments}
We would like to thank Lorenzo Cornalba,
Steve Gubser, Dan Kabat, Igor Klebanov, Morten Krogh and Ricardo Schiappa for
helpful conversations. The work of MVR is supported in part by the
Natural Sciences and Engineering Research Council of Canada (NSERC).
The work of WT is supported in part by the A.\ P.\ Sloan Foundation
and in part by the DOE through contract \#DE-FC02-94ER40818.
|
\section{Introduction}
\setcounter{equation}{0} The problem of integrability and nonintegrability
of Hamiltonian system with $g$ degrees of freedom has been a subject of
considerable interest for many years. Recently remarkable progress has been
achieved in connection of study of stationary soliton type equations. The
method of restricted flows was introduced in \cite{Mos1,20,Mos2} as a
non-linearization of the Korteweg-de Vries (KdV) spectral problem and was
generalized in \cite{5,2}. The coupled Neumann system, the Neumann system,
the Garnier type systems, the Rosochatius-Wojciechowski, Rosochatius and the
H\'enon-Heiles type
systems are examples of this type. In this relation important results about
the algebro-geometrical interpretaion of these systems are obtained in \cite
{F}. General approach to completely integrable dynamical systems of Neumann
type is discussed in \cite{S, AHP, AHH}. Quasi-periodic solutions and
spectral interpretation using both algebro-geometrical and spectral methods
are given in \cite{K}. Quasi-periodic solutions ($g=2$) and
periodic solutions associated to Lam\'e and Treibich-Verdier potentials are
obtained in \cite{EK, CEEK}. New method of constructing elliptic finite-gap
solutions of the stationary KdV hierarchy, based on a theorem due to Picard,
is proposed in \cite{GW95a,GW95b,GW95c,GW95d,GW96,GW98}.
In this paper we are concerned with the following different approaches
\begin{itemize}
\item integrable dynamical systems related to Hill's equation in the case
of finite-gap potential \cite{20, AM80, F, AHP, AHH, S, K, CC}
\item method of stationary flows and restricted flows \cite{1,2}
\item method of non-linearization of the KdV spectral problem \cite{5}
\item method of separation of variables \cite{KKM, KRT, EEKT}
\item algebro-geometrical construction \cite{F,S,K}.
\end{itemize}
We give unified construction based on algebro-geometrical approach. New
solutions in terms of Novikov and Gelfand-Dickey (GD) polynomials are given
in explicit form.
The paper is organized as follows. In section 2 we construct the stationary
flows associated to the KdV hierarchy strictly following \cite{1,7,8,9}.
Some maps between completely integrable dynamical systems are presented. In
section 3 we analyse the relation between restricted KdV flows and Garnier
type systems following \cite{8}. The map between stationary and restricted
KdV flows are given in section 4 \cite{8}. The list of known dynamical
systems related to stationary KdV equations are presented in section 3.
In section 4 we formulate Baker-Akhiezer function approach from
algebro-geometrical and spectral point of view respectively. New solutions
of integrable dynamical system associated to stationary KdV equations in
terms of Novikov polynomials are presented . In
particular these polynomials associated to Lam\'e potentials coincide with
Hermite polynomials and Lam\'e polynomials. In section 5 ($2\times 2$) Lax
representations of Garnier, Rosochtius I and stationary second flow of
vector nonlinear Schr\"{o}dinger equation are given. Following \cite
{EEKT,KRT,T1} $r$-matrix approach of these systems is discussed.
\section{KdV hierarchy and Gelfand-Dickey polynomials}
\setcounter{equation}{0} In this section we follow the geometrical
construction of \cite{1,7,8,9,v98} with only little changes of signs and
different spectral parameter. Let $M$ be a bi-Hamiltonian manifold: if the
associated Poisson operator $P^{\lambda}:=P_{1}+4\lambda P_{0}$ admits as a
Casimir a formal Laurent series $h(\lambda)$
\begin{equation}
h(\lambda):=\sum_{j\geq0}\,h_{j}\lambda^{-j}
\end{equation}
then $h_{0}$ is a Casimir of $P_{0}$ and the coefficients $h_{j}$ ($j\geq0$)
are the Hamiltonian functions of a hierarchy of bi-Hamiltonian vector fields
$X_{j}$:
\begin{equation}
X_{j}=P_{1}dh_{j}=P_{0}dh_{j+1}, \quad (j\geq0) \label{VecF}
\end{equation}
At any point $u\in M$, the bi-Hamiltonian flows are given by $\frac{d u}{%
dt_{j}} =X_{j}(u)$, $t_{j}$ being the evolution parameter of the $j$th flow.
The vector field (\ref{VecF}) are Hamiltonian also with respect to the
Poisson operator $P^{\lambda}$. In fact the recursion relation (\ref{VecF})
can be written as
\begin{equation}
X_{j}=P^{\lambda}\,dh^{(j)}(\lambda),\quad h^{(j)}(\lambda):=(\lambda^{j}
h(\lambda))_{+} \label{X}
\end{equation}
where the index $+$ means the projection of a Laurent series onto the purely
polynomial part.
Let $M$ be the algebra of polynomials in $u, u_{x},u_{xx},\ldots$ ($u=u(x)$
is a $C^{\infty}$ function of $x$ and the subscript $x$ means the derivative
with respect to $x$), and le $P_{0}$ and $P_{1}$ be the two Poisson
operators of the KdV hierarchy \cite{11}:
\begin{equation}
P_{0}:=\frac{d}{dx}, \qquad P_{1}:=\frac{d^{3}}{dx^{3}} - 4 u \frac{d}{dx} -
2 u_{x} . \label{Opers}
\end{equation}
The gradients of the Casimirs of the associated Poisson operator $%
P^{\lambda} $ can be obtained searching the functions $v(\lambda):=\sum_{j%
\geq0}v_{j}\lambda^{-j}$ which are solutions of the following equation:
\begin{equation}
B^{\lambda}(v(\lambda),v(\lambda))=a(\lambda) \label{GDg}
\end{equation}
where $a(\lambda)=\sum_{j\geq0}a_{j}\lambda^{-j}$, $a_{j}$ are constant
parameters and $B^{\lambda}$ is the bilinear function
\begin{equation}
B^{\lambda}(w_{1},w_{2}):=w_{1xx}w_{2}+w_{1}w_{2xx}-w_{1x}w_{2x}-
4(u-\lambda)w_{1}w_{2} . \label{Brel}
\end{equation}
In fact $B^{\lambda}$ is related to the Poisson operator through the
relation
\begin{equation}
\frac{d}{dx}B^{\lambda}(w_{1},w_{2})=w_{1} P^{\lambda} w_{2} +w_{2}
P^{\lambda} w_{1} .
\end{equation}
Equation (\ref{GDg}) can be solved developing the left-hand side as a
Laurent series
\begin{equation}
B^{\lambda}(v(\lambda),v(\lambda)) = \sum_{k\geq -1} B_{k}\lambda^{-k}
\end{equation}
so that, for each $a(\lambda)$, it furnishes the coefficient of the solution
$v(\lambda)$ (unique up to a sign). The solution corresponding to $\bar{a}%
(\lambda) =-\lambda$ is the so-called basis solution $\bar{v}(\lambda)$; its
first coefficients are
\begin{eqnarray}
\bar{v}_{0}=1,\quad \bar{v}_{1}=1, \quad \bar{v}=2(u_{xx}+3 u^{2}), \\
\bar{v}_{3}=2(u_{xxxx}+5u_{x}^{2}+10u_{xx}u+10u^{3}) ,
\end{eqnarray}
and so on, namely the gradients of the first KdV Hamiltonians. In what
follows we shall consider also the function $v(\lambda)=c(\lambda)\bar{v}%
(\lambda)$, which is a solution of (\ref{GDg}) for
\begin{equation}
a(\lambda)=-\lambda c^{2}(\lambda), \quad c(\lambda)=1 + \sum_{j\geq
1}\,c_{j}\lambda^{-j} ,
\end{equation}
where the coefficients $c_{j}$ are free parameters. In this case the first
1-forms of the hierarchy are $v_{0}=1, v_{1}=\bar{v}_{1}+c_{1}, v_{2}=\bar{v}%
_{2}+c_{1}\bar{v}_{1}+c_{2}$, and so on.
The coefficient $B_{k}$ can be expressed through the GD polynomials. For
each Laurent series $v(\lambda)$ let us consider the functions $%
B^{(k)}(\lambda):=B^{\lambda}(v(\lambda),v^{(k)}(\lambda))$, where $%
v^{(k)}(\lambda):=\left(\lambda^{k}v(\lambda)\right)_{+}$; these functions
have the form
\begin{equation}
B^{(k)}(\lambda)=\lambda^{k+1}v_{0}^{2}+
\sum_{j=1}^{k-1}\lambda^{k-j}(p_{0j} +v_{0}v_{j+1}) +\sum_{j\geq
0}\lambda^{-j}p_{jk} . \label{BB}
\end{equation}
It can be shown that
\begin{equation}
B_{-1}=-v_{0}^{2}, \quad B_{k}=p_{0k}-v_{0}v_{k+1} .
\end{equation}
Furthermore, if $v(\lambda)$ is a solution of (\ref{GDg}), the coefficients $%
p_{jk}$ in (\ref{BB}) are polynomials in $u$ and its $x$-derivatives. They
will be referred to as Gelfand-Dickey (GD) polynomials and the function $%
B^{\lambda}$ as their generating function.
The fundamental property of the GD polynomials, stemming from (\ref{GDg}), (%
\ref{BB}) is the following relation with the gradients $v_{j}=dh_{j}$ and
the bi-Hamiltonian vector field $X_{k}$:
\begin{equation}
\frac{d}{dx}p_{jk}=v_{j}X_{k} .
\end{equation}
We report some GD polynomials to be used in what follows $(v_{0}=1)$:
\begin{eqnarray} \label{GD}
p_{00}&=&4 u - v_{1} , \nonumber \\
p_{01}&=&8 u v_{1}-v_{1}^{2}-v_{2}+2 v_{1xx} , \nonumber \\
p_{02}&=&4uv_{1}^{2}+8uv_{2}-2v_{1}v_{2}-v_{3}-v_{1x}^{2}+
2v_{1}v_{1xx}+2v_{2xx}, \nonumber \\
\\
p_{12}&=&8uv_{1}v_{2}-v_{2}^{2}+4uv_{3}-v_{1}v_{3}-v_{4}+
2v_{1x}v_{2x}+2v_{2}v_{1xx}+2v_{1}v_{2xx}+v_{3xx} , \nonumber \\
p_{kk}&=&2v_{kxx}v_{k}-v_{k}^{2}+4uv_{k}^{2} . \nonumber
\end{eqnarray}
The GD polynomials corresponding to the basis solution $\bar{v}(\lambda)$
are the polynomials defined in \cite{1}, proposition 12.1.12.
\subsection{The method of stationary flows}
The method of stationary flows was developed in
order to reduce the flows of the KdV hierarchy onto the set $M_{g}$ of fixed
points of the $g$th flow $X_{g}$ of the hierarchy
\begin{equation}
M_{g}:=\left\{u|X_{g}(u,u_{x},\ldots,u^{(2g+1)})=0\right\}.
\end{equation}
As $M_{g}$ is odd-dimensional it can not be a symplectic manifold;
nevertheless we will show that it is a bi-Hamiltonian manifold; it will be
referred to as extended phase space. Moreover, $M_{g}$ is naturally
foliated, on account of (\ref{VecF}) and (\ref{Opers}), by a one-parameter
family of $2g$-dimensional submanifolds $S_{g}$ given by
\begin{equation}
S_{g}:=\left\{u|v_{g+1}(u,u_{x},\ldots,u^{(2g)}=c\right\}
\end{equation}
($c$ being a constant parameter), which are invariant manifolds with respect
to each vector field of the KdV hierarchy, due to the invariance of the
functions $v_{k}$. So $M_{g}$ can be parametrized naturally by $%
v_{1},\ldots, v_{g+1}$ and by their $x$-derivatives $v_{1x},\ldots,v_{gx}$.
We shall use these coordinates in what follows.
From the computational point of view, one proceeds as follows.
(i) Due to (\ref{X}) and (\ref{GDg}), the manifold $M_{g}$ is defined by the
solutions $u$ of the equation
\begin{equation}
B^{\lambda}(v(\lambda),v^{(g)}(\lambda))=\lambda^{g}a(\lambda) \label{BL}
\end{equation}
where $v(\lambda)=\sum_{j=0}^{g}v_{j}\lambda^{-j}$, $a(\lambda)=%
\sum_{j=-1}^{2g} a_{j}\lambda^{-j}$. In particular, if $a(\lambda)=-\lambda
c^{2}(\lambda)$, $M_{g}$ is given by
\begin{equation}
M_{n}=\left\{u|\bar{X}_{n}+\sum_{j=1}^{n}c_{j}\bar{X}_{n-j} = 0\right\}
\end{equation}
i.e. by the solutions of the Lax-Novikov equations. Taking into acount (\ref
{BB}) and choosing $a_{-1}=-1$, by equation (\ref{BL}) the coefficients of $%
\lambda^{g+1}$ we get $v_{0}^{2}=1$; from now on we put $v_{0}=1$. Moreover,
equating the coefficients of the other powers of $\lambda$ we get the
following system:
\begin{equation}
p_{0k} - v_{k+1} =a_{k},\quad (k=0\ldots, n-1)\quad p_{jn}=a_{n+j},\,
(j=0,\ldots,n). \label{PP0}
\end{equation}
(ii) In order to obtain the first Poisson tensor $P_{0}$, we eliminate $%
u=v_{1}/2+a_{0}/4$ from (\ref{PP0}) using the first equation ($k=0$) and we
extract the system of $n$ second-order ODEs in the $v_{j}$ ($j=1,\ldots,n$):
\begin{equation}
p_{0k} - v_{k+1} =a_{k},\quad (k=1\ldots, g-1)\quad p_{0g}=a_{g} \label{P0}
\end{equation}
which will be referred as $P_{0}$-system. The remaining equations (\ref{PP0}%
) will furnish a set of $g$ independent integrals of motion. In order to
obtain a second Poisson structure, we consider the following system: ($P_{1}$%
-system)
\begin{equation}
p_{0k} - v_{k+1} =a_{k},\quad (k=1\ldots, g-1)\quad p_{gg}=a_{2g},
\label{P1}
\end{equation}
with $u$ as above.
(iii) The $P_{0}$-system (\ref{P0}) and the $P_{1}$-system (\ref{P1}) can be
written as canonical Hamiltonian systems
\begin{equation}
r_{kx}=\frac{\partial H_{g}^{(0)}}{\partial s_{k}}, \qquad s_{kx}=-\frac{%
\partial H_{g}^{(0)}}{\partial r_{k}},
\end{equation}
\begin{equation}
q_{kx}=\frac{\partial H_{g}^{(1)}}{\partial p_{k}}, \qquad p_{kx}=-\frac{%
\partial H_{g}^{(1)}}{\partial q_{k}} \label{Deq}
\end{equation}
have $n$ integrals of motion given by
\begin{eqnarray}
&&K_{j}=-\frac{1}{8}p_{jg}|_{r}=a_{g+j}, (j=1,\ldots,g), \\
&&H_{j}=-\frac{1}{8}p_{jg}|_{x}=a_{g+j}, (j=0,\ldots,g-1),
\end{eqnarray}
Moreover, the map $\Phi:M_{g}\rightarrow M_{g}$ in the extended phase space
generates a second Poisson structure.
\section{Integrable dynamical systems related to hierarchy of stationary KdV
equations}
\setcounter{equation}{0} In the recent years, remarkable progress has been
achieved in the description of those quasi(periodic) potentials which belong
to a given spectrum. Many integrable systems of differential equations are
shown to be closely connected with Hill's equation in the case of a finite
gap potential. The coupled Neumann system, The Neumann system, and the
Rosochatius systems are examples of this type.
In this paragraph we are concerned with the following completely integrable
systems
The Garnier system
\begin{eqnarray}
\xi_{ixx} = \left( 2\sum_{j=1}^{g} \xi_{j} \eta_{j} +
\tilde{a}_{i}\right) \xi_{i}
, \quad \eta_{ixx} = \left( 2\sum_{j=1}^{g} \xi_{j} \eta_{j} +
\tilde{a}_{i}\right) \eta_{i} . \label{Garn}
\end{eqnarray}
The $g$-dimensional anisotropic harmonic oscillator in radial quartic
potential, is obtained when $\xi_{i}=\eta_{i}$ $i=1,\ldots, g$
\begin{eqnarray}
\xi_{ixx} = \left( 2\sum_{j=1}^{g} \xi_{j}^2 + \tilde{a}_{i}\right) \xi_{i} .
\label{oscil}
\end{eqnarray}
Another interesting integrable system was proposed recently \cite{W}, we
call it the Rosochatius-Wojciechowski system. In our context, this system is
obtained by the Deift elimination procedure. Let
\begin{equation}
\xi_{i}=\psi_{i}\exp(\theta_{i}),\, \eta_{i}=\psi_{i}\exp(-\theta_{i}) ,\,
\sqrt{f}_{i} = \psi_{i}^{2}\theta_{ix} , \label{Deift}
\end{equation}
then equations (\ref{Garn}) transform to the Rosochatius-Wojciechowski system
\begin{eqnarray}
\psi_{ixx} = \left( 2\sum_{j=1}^{g} \psi_{j}^2 \right) \psi_{i}
+\tilde{a}_{i}\psi_{i}-f_{i}/\psi_{i}^3 , \,i=1,\ldots ,g \label{RosI}
\end{eqnarray}
and the Hamiltonian is given by
\begin{eqnarray}
H=\sum_{j=1}^{g}\,\chi_{j}^{2} -
\left(\sum_{k=1}^{g}\,\psi_{k}^{2}\right)^{2} - \sum_{j=1}^{n}
\tilde{a}_{j}\psi_{j}^{2} -\sum_{j=1}^{g} \frac{f_{j}}{\psi_{j}^{2}} ,
\end{eqnarray}
where $\chi_{j}=\psi_{jx}$ are canonical momenta.
(ii) The coupled Neumann system
\begin{eqnarray}
\tilde{\xi}_{ixx} + \left( 2\sum_{j=0}^{g} b_{j}\tilde{\xi}_{j} \tilde{\eta}%
_{j} +\tilde{\xi}_{jx}\tilde{\eta}_{jx}\right)\tilde{\psi}_{i}= b_{i} \tilde{%
\xi}_{i} , \\
\tilde{\eta}_{ixx} + \left( 2\sum_{j=0}^{g} b_{j}\tilde{\xi}_{j} \tilde{\eta}%
_{j} +\tilde{\xi}_{jx}\tilde{\eta}_{jx}\right)\tilde{\eta}_{i}= b_{i} \tilde{%
\eta}_{i} . \label{CouplNeum}
\end{eqnarray}
with constraint $\sum_{i=0}^{g}\tilde{\xi}_{i}\tilde{\eta}_{i} =1$, where $%
b_{0}< b_{1}< \ldots < b_{g}$ are fixed real numbers. The Neumann system is
obtained when $\xi_{i}=\eta_{i}$
\begin{eqnarray}
\tilde{\xi}_{ixx} + \left( 2\sum_{j=0}^{g} b_{j}\tilde{\xi}_{j}^2 +\tilde{\xi%
}_{jx}^2 \right)\tilde{\xi}_{i}= b_{i} \tilde{\xi}_{i} . \label{Neum}
\end{eqnarray}
This system describes the motion of uncoupled harmonic oscillators $\tilde{%
\xi}_{ixx} = b_{i} \tilde{\xi}_{i}$, constrained by the force $%
\sum_{i=0}^{g}(b_{i}\tilde{\xi}_{i}^2 +\tilde{\xi}_{ix}^2)$ to move on the
unit sphere $\sum_{i=0}^{g}\tilde{\xi}_{i}^2=1$.
Let
\begin{equation}
\tilde{\xi}_{i}=\tilde{\psi}_{i}\exp(\tilde{\theta}_{i}),\,
\tilde{\eta}_{i}=\tilde{\psi}_{i}\exp(-\tilde{\theta}_{i}) ,\,
\sqrt{\tilde{f}}_{i} = \tilde{\psi}_{i}^{2}\tilde{\theta}_{ix} ,
\label{Deift1}
\end{equation}
then by Deift procedure (\ref{Deift1}), equations (\ref{CouplNeum}) transform
to Rosochatius system
\begin{eqnarray}
\tilde{r}_{ixx} = -\left( \sum_{j=0}^{g} b_{j}\tilde{r}_{j}^2 + \
\tilde{r}_{jx}^2- \frac{\tilde{f}_{j}}{\tilde{r}_{j}^{2}}
\right)\tilde{r}_{i}-\frac{\tilde{f}_{j}}{\tilde{r}_{j}^{3}}+
b_{i} \tilde{r}_{i} . \label{Ros}
\end{eqnarray}
where $\sum_{i=0}^{g}\tilde{r}_{i}^2=1$.
\section{Baker-Akhiezer function}
\setcounter{equation}{0}
We review in this section some basic facts about Baker--Akhieser function
which will be used in the sequel.
Let $K$ be the hyperelliptic Riemann surface $\mu^2=\prod_{i=0}^{2g}
(\lambda-\lambda_{i})=R(\lambda ) $. The points of $K$ are pairs $%
P=(\lambda,R)$ and $\lambda(P)$ is the value of the natural projection $%
P\rightarrow\lambda(P)$ of $K$ to the complex projective line $CP^{1}$.
For given nonspecial divisor $D$, there is an unuque Baker-Akhiezer(BA)
function $\Psi(t,\lambda)$, such that
(i) the divisor of the poles of $\Psi$ is $D$,
(ii) $\Psi$ is meromorphic on $K\backslash\infty$
(iii) when $P\rightarrow \infty$
\begin{eqnarray}
\Psi(x,P)\exp(-kx) = 1 +\sum_{s=1}^{\infty} m_{s}(x) k^{-s} , \label{expan1}
\end{eqnarray}
is holomorphic and $k=\sqrt{\lambda(P)}$ is a local parameter near $P=\infty$%
.
There is a unique function $u(x)$ such that
\begin{equation}
\Psi_{xx} - u(x)\Psi= \lambda(P)\Psi , \label{Hill}
\end{equation}
where $\Psi$ is a BA function. Inserting expansion (\ref{expan1}) into (\ref
{Hill}), we obtain
\begin{eqnarray}
\Psi_{xx} - 2 m_{1x}(x)\Psi -\lambda(P)\Psi =\exp(kx) O(k^{-1}) ,
\end{eqnarray}
and due to the uniqueness of $\Psi$, we prove (\ref{Hill}), with $u(x)=2
m_{1x}(x)$.
By the Riemann-Roch theorem, there exists a unique differential $\tilde{%
\Omega}$ and a nonspecial divisor $D^{\tau}$ of degree $g$ such that the
zeros of $\tilde{\Omega}$ are $D + D^{\tau}$ and the expansion at $P=\infty$%
, $\tilde{\Omega}(P)=(1+O(k^{-2})) dk$.
For given nonspecial divisor $D^{\tau}$, there exists a unuque dual
Baker-Akhiezer (BA) function such that
(i) the divisor of the poles of $\Psi$ is $D^{\tau}$,
(ii) $\Psi$ is meromorphic on $K\backslash\infty$
(iii) when $P\rightarrow \infty$
\begin{eqnarray}
\Psi^{\tau}(x,P)\exp(-kx) = 1 +\sum_{s=1}^{\infty}\tilde{m}_{s}(x) k^{-s},
\label{expan}
\end{eqnarray}
Fix $\tau$ to be the hyperelliptic involution $P=(\lambda,R)\rightarrow(%
\lambda,-R)$, then we have $D^{\tau}=\tau D$, $\Psi^{\tau}(x,P)=\Psi(x,\tau
P)$. Let $\sum_{i=1}^{g}\mu_{i}(0)$ be the $\lambda$-projection of $D$, and $%
\sum_{i=1}^{g}\mu_{i}(x)$ be the $\lambda$-projection of the zero divisor of
$\Psi(x,P)$. The function $\Psi(t,P)\Psi^{\tau}(t,P)$ is meromorphic on $%
{\bf CP^{1}}$ and the following identity takes place
\begin{equation}
\Psi(x,P)\Psi^{\tau}(x,P)=\frac{F(x,\lambda)}{F(0,\lambda)} \label{sqfun}
\end{equation}
where $F(x,\lambda)=\prod_{i=1}^{g}(\lambda - \mu_{i}(x))$. Introduce the
Wronskian
\begin{eqnarray}
\left\{\Psi(x,P),\Psi^{\tau}(x,P)\right\}=&&\Psi_{x}(x,P)\Psi^{\tau}(x,P)
-\Psi(x,P)\Psi_{x}^{\tau}(x,P) = \\
&&\frac{2\sqrt{R(\lambda)}}{\prod_{i=1}^{g}(\lambda -\mu_{i}(0)} , \nonumber
\end{eqnarray}
and the differential $\tilde{\Omega}$ is given explicitly by
\begin{equation}
\tilde{\Omega}(P)=\frac{1}{2}\prod_{i=1}^{g} (\lambda - \mu_{i}(0))/\sqrt{%
R(\lambda)}d\lambda . \label{diff}
\end{equation}
We assume that $E(P)$ is a meromorphic function on $K$ with $g+1$ simple
poles $\infty,p_{1},\ldots,p_{g}$ and at $P\rightarrow\infty$, $%
E(P)=k+\ldots $, and $\tilde{E}(P)$ is meromorphic function with $g+1$
simple poles $q_{0},q_{1},\ldots ,q_{g}$ and at $P\rightarrow\infty$, $%
\tilde{E}(P)=k^{-1} +\ldots$. We also suppose that the divisors of poles of $%
E(P)$ and $\tilde{E}(P)$ are different from $D$, $D^{\tau}$.
Let
\begin{eqnarray}
\tilde{\xi}_{i}=\tilde{\xi}_{i}^{0}\Psi(x,q_{i}),\quad \tilde{\eta}_{i}=%
\tilde{\eta}_{i}^{0}\Psi^{\tau}(x,q_{i}),\quad \nonumber \\
\tilde{\xi}^{0}_{i}\tilde{\eta}_{i}^{0}=Res_{P=q_i}E\tilde{\Omega}, \quad
b_{i}=\lambda(q_{i}),\,i=0,\ldots ,g \label{solu1} \\
\xi_{i}=\xi_{i}^{0}\Psi(x,p_{i}),\quad
\eta_{i}=\eta_{i}^{0}\Psi^{\tau}(x,p_{i}),\quad \nonumber \\
\xi^{0}_{i}\eta_{i}^{0}=Res_{P=q_i}\tilde{E}\tilde{\Omega}, \quad
a_{i}=\lambda(p_{i}),\,i=1,\ldots ,g \label{solu2}
\end{eqnarray}
then
\begin{eqnarray}
& &u(x)=-\left(\sum_{i=0}^{g}b_{i}\tilde{\xi}_{i}\tilde{\eta}_{i} +\tilde{\xi%
}_{ix}\tilde{\eta}_{ix}\right),\quad \sum_{i=0}^{g}\tilde{\xi}_{i}\tilde{\eta%
}_{i}=1 \nonumber \\
& &u(x)=2\sum_{i=1}^{g}\xi_{i}\eta_{i}+ const. \label{poten}
\end{eqnarray}
Let us construct the meromorphic differential $\tilde{E}\Psi\Psi^{\tau}%
\tilde{\Omega}$. By direct computations we have
\begin{equation}
\sum_{i=0}^{g}\,Res_{P=q_{i}}\,\tilde{E}\Psi\Psi^{\tau}\tilde{\Omega}
+Res_{P=\infty}\,\tilde{E}\Psi\Psi^{\tau}\tilde{\Omega} =\sum_{i=0}^{g}%
\tilde{\xi}_{i}\tilde{\eta}_{i} - 1 = 0,
\end{equation}
where $\tilde{\xi}_{i}^{0}\tilde{\eta}_{i}^{0}=Res_{P=q_{i}}\,\tilde{E}
\tilde{\Omega}$. Differentiating $\sum_{i=0}^{g}\tilde{\xi}_{i}\tilde{\eta}%
_{i} = 1$ twice and using Eq. (\ref{solu1}), we obtain first expression in (%
\ref{poten}). The eigenvalue equations
\begin{eqnarray}
&&\tilde{\xi}_{ixx}=(\lambda(q_{i}) + u(x))\tilde{\xi}_{i}, \\
&&\tilde{\eta}_{ixx}=(\lambda(q_{i}) + u(x))\tilde{\eta}_{i},
\end{eqnarray}
by replacing $u(x)$ from (\ref{poten}) are the coupled Neumann system (\ref
{CouplNeum}). By computations of the same kind, we have
\begin{equation}
\sum_{i=1}^{g}\,Res_{P=p_{i}}\,E\Psi\Psi^{\tau}\tilde{\Omega}
+Res_{P=\infty}\,E\Psi\Psi^{\tau}\tilde{\Omega} =\sum_{i=1}^{g} \xi_{i}
\eta_{i} - u(x)+\frac{1}{2} const. = 0,
\end{equation}
where $\xi_{i}^{0} \eta_{i}^{0}=Res_{P=p_{i}}\, E\tilde{\Omega}$. The
corresponding eigenvalue equations are the Garnier system (\ref{Garn}).
\subsection{Spectral interpretation}
Let $p$ be a positive real divisor of degree $g$ on a real hyperelliptic
curve
\begin{equation}
\mu^{2}=R(\lambda)=\prod_{i=0}^{2g}(\lambda-\lambda_{i}),\quad \lambda_{0} <
\lambda_{1} <,\ldots ,<\lambda_{2g} .
\end{equation}
The projection $\lambda(p_{i})$ lie in the closed lacunae $%
[\lambda_{2i-1},\lambda_{2i}]$. The following considerations, due to Jacobi,
can be used to construct such a divisor.
Each divisor $p$ determines and is determined by a system of polynomials
\begin{eqnarray}
&&\tilde{A}(\lambda)=\prod_{i=1}^{g}(\lambda-\lambda(p_{i})),\quad \tilde{C}%
(\lambda)=\tilde{A}(\lambda)\, \sum_{i=1}^{g}\,\frac{\sqrt{R(p_{i})}} {%
\tilde{A}^{\prime}(p_{i})(\lambda-\lambda(p_{i}))} , \nonumber \\
&&\tilde{B}(\lambda)=\lambda^{g+1}+\ldots ,
\end{eqnarray}
of degrees $g$, $g-1$, $g+1$, respectively, with $R=\tilde{C}^2-\tilde{A}%
\tilde{B}$. The complementary divisor $q$ is also determied by this
construction. This is the content of the following step.
For a given spectral data
\begin{equation}
\lambda_{0}=0 < \lambda_{1} < ,\ldots ,< \lambda_{2g},\quad
\lambda(p_{i})\in [\lambda_{2i-1},\lambda_{2i}],\,i=1,\ldots,g
\end{equation}
thete exists
\begin{equation}
\lambda(q_{i}),\,\,i=0,\ldots,g\quad\lambda(q_{0})\in
(-\infty,\lambda_{0}],\quad \lambda(q_{i})\in[\lambda_{2i-1},\lambda_{2i}] ,
\end{equation}
such that $R=\tilde{C}^2-\tilde{A}\tilde{B}$, and the projections $%
\lambda(q_{i})$ are the roots of $\tilde{B}$.
Note that the functions $E(P)$, $\tilde{E}(P)$ are meromorphic on $K$ and
the following formulas are immediate
\begin{equation}
E(P)=(\sqrt{R(\lambda)} +\tilde{C}(\lambda))/\tilde{A}(\lambda), \qquad
\tilde{C}(p_{i})=\sqrt{R(p_{i})} \label{E}
\end{equation}
\begin{equation}
\tilde{E}(P)=(\sqrt{R(\lambda)} +\tilde{C}(\lambda))/\tilde{B}(\lambda),
\qquad \tilde{C}(q_{i})=-\sqrt{R(q_{i})} \label{tE}
\end{equation}
Now we recall some facts from the periodic theory of Hill's equation. We
suppose that $u(x)$ is a real finite-gap potential, i.e. the operator $L$
has only $2g+1$ simple eigenvalues $\lambda_{0} < \lambda_{1} < ,\ldots
,<\lambda_{2g}$ and the rest of the spectrum consists of double eigenvalues.
The periodic spectra of $L$ is determined by the combined eigenvalues of the
periodic
\begin{equation}
L\,f_{2i} = \lambda_{2i}\,f_{2i},\quad f(x+1)=f(x),\,i=0,\ldots,g
\end{equation}
and the antiperiodic
\begin{equation}
L\,f_{2i-1} = \lambda_{2i-1}\,f_{2i-1},\quad f(x+1)=-f(x),\,i=1,\ldots,g
\end{equation}
eigenvalue equations. The intervals $(-\infty,\lambda_{0}],[\lambda_{2i-1},%
\lambda_{2i}]$ are termed lacunae. The Floquet solutions(periodic BA
function) and the corresponding Floquet multipliers, are given by
\begin{eqnarray}
&&\Psi(x,\lambda) =\left[ F(x,\lambda)/F(0,\lambda)\right]^{1/2}\,
\exp\left(\int_{0}^{x}\,\sqrt{R(\lambda)}/F(x^{\prime},\lambda)dx^{\prime}%
\right) , \label{BA1} \\
&&\Psi(x+1,\lambda)=\rho_{+}(\lambda)\Psi(x,\lambda) , \\
&&\Psi^{\tau}(x,\lambda) =\left[ F(x,\lambda)/F(0,\lambda)\right]^{1/2}\,
\exp\left(-\int_{0}^{x}\,\sqrt{R(\lambda)}/F(x^{\prime},\lambda)dx^{\prime}%
\right) , \label{BA2} \\
&&\Psi^{\tau}(x+1,\lambda)=\rho_{-}(\lambda)\Psi^{\tau}(x,\lambda) , \qquad
\rho_{\pm}=\exp(\pm\tilde{p}(\lambda)) ,
\end{eqnarray}
where
\begin{equation}
\tilde{p}(\lambda)=\int_{0}^{1}\,\sqrt{R(\lambda)}/F(x,\lambda)dx
\end{equation}
Note that if $\lambda$ is in the periodic spectrum, $\Psi(x,%
\lambda_{2i})=f_{2i}$, $i=0,\ldots,g$ is a periodic eigenfunction, and $%
\Psi(x,\lambda_{2i-1})=f_{2i-1}$,$i=1,\ldots ,g$ is an antiperiodic
eigenfunction. It is well known that the projections of the zeros of the
Floquet solution define the auxiliary spectrum of $L$.
The following expressions hold
\begin{eqnarray}
&& \xi_{i} \eta_{i}=\prod_{j=1}^{g}\,(\lambda(p_{i})-\mu_{j}(x))/\tilde{A}%
^{\prime}(p_{i}), \label{Sol1} \\
&& \xi_{i}^{0} \eta_{i}^{0}=\prod_{j=1}^{g}\,(\lambda(p_{i})-\mu_{j}(0))/
\tilde{A}^{\prime}(p_{i}), \nonumber \\
&& \tilde{\xi}_{i}\tilde{\eta}_{i}=\prod_{j=1}^{g}\,(\lambda(q_{i})-%
\mu_{j}(x))/ \tilde{B}^{\prime}(q_{i}), \label{Sol2} \\
&& \tilde{\xi}_{i}^{0} \tilde{\eta}_{i}^{0}=\prod_{j=1}^{g}\,(%
\lambda(q_{i})-\mu_{j}(0))/ \tilde{B}^{\prime}(q_{i}). \nonumber
\end{eqnarray}
Using (\ref{E}),(\ref{tE}) we obtain
\begin{eqnarray}
Res_{P=p_{i}}\,E(P)\Psi\Psi^{\tau}\tilde{\Omega}& & =
x_{i}^{0}y_{i}^{0}\Psi(x,p_{i})\Psi^{\tau}(x,p_{i}) \nonumber \\
& &=\prod_{j=1}^{g}\,(\lambda(p_{i})-\mu_{j}(x))/A^{\prime}(p_{i}) ,
\nonumber
\end{eqnarray}
where $\xi_{i}^{0}\eta_{i}^{0}$ is given by (\ref{Sol1}).
Let
\begin{eqnarray}
e_{2i}^2=\frac{\prod_{i\neq j}^{g}\,(\lambda_{2i} -\lambda_{2j})}
{\prod_{j\neq j=0}^{g}\,(\lambda_{2i}-\mu_{j}(0))} ,\quad f_{2i}^2=\frac{%
\prod_{j=1}^{g}\,(\lambda_{2i} -\mu_{j}(t))} {\prod_{i\neq
j=0}^{g}\,(\lambda_{2i}-\mu_{j}(0)) } = \Psi^2(\lambda_{2i}) ,
\end{eqnarray}
$i=0,\ldots ,g$, then the expressions (\ref{poten}), (\ref{Sol2}) are the
famous McKean-Moerbeke expansion of the potential $u(x)$ in terms of squares
of the eigenfunctions
\begin{equation}
u(x)=-\left( 2\sum_{i=0}^{g}\,\lambda_{2i} f_{2i}^2/e_{2i}
+\sum_{i=1}^{g}\lambda_{2i-1}-\sum_{i=1}^{g}\lambda_{2i} +\lambda_{0}\right),
\end{equation}
where the following identity among the squares of eigenfunctions hold on
\begin{equation}
\sum_{i=0}^{g}\,e_{2i}^{-2} f_{2i}^{2} =1.
\end{equation}
The results of this section, may be summarized by following:
Let $u(x)$ be a real nonsingular finite-gap potential. There exists $g$
eigenfunctions $\Psi(p_{1}),\ldots ,\Psi(p_{g})$ and $g+1$ eigenfunctions $%
\Psi(q_{0}),\ldots ,\Psi(q_{g})$ of Hill's equation, corresponding to the
eigenvalues $\lambda(p_{1}),\ldots,\lambda(p_{g})$ and $\lambda(q_{0}),%
\ldots ,\lambda(q_{g})$, respectively, such that
(i)
\begin{eqnarray}
u(x) = 2\sum_{i=1}^{g}\,\Psi(p_{i})\Psi^{\tau}(p_{i}) e_{i}^{-2} +
2\sum_{i=1}^{g}\lambda(p_{i})-\sum_{i=0}^{2g}\lambda_{i} , \\
e_{i}^{-2}=\prod_{j=1}^{g}(\lambda(p_{i})-\mu_{j}(0))/\tilde{A}%
^{\prime}\quad \Psi(p_{i})\equiv\Psi(x,\lambda)|_{\lambda =p_{i}} ,\,\,
i=1,\ldots,g . \nonumber
\end{eqnarray}
(ii)
\begin{eqnarray}
u(x) = 2\sum_{i=0}^{g}\,\lambda(q_{i})\Psi(q_{i})\Psi^{\tau}(q_{i})
e_{i}^{-2} - 2\sum_{i=0}^{g}\lambda(q_{i})+\sum_{i=0}^{2g}\lambda_{i} ,
\label{Expan1} \\
\tilde{e}_{i}^{-2}=\prod_{j=1}^{g}(\lambda(q_{i})-\mu_{j}(0))/\tilde{B}%
^{\prime}, \,\, \Psi(q_{i})\equiv\Psi(t,\lambda)|_{\lambda =q_{i}} ,\,\,
i=0,\ldots,g , \nonumber \\
\sum_{i=0}^{g}\tilde{e}^{-2}\Psi(q_{i})\Psi^{\tau}(q_{i}) =1 .
\end{eqnarray}
The corresponding eigenvalue equations are the Garnier and coupled Neumann
system.
Let
\begin{eqnarray}
& &e_{2i-1}^{2}=\prod_{i\neq j}^{g}(\lambda_{2i-1} - \lambda_{2j-1})/
\prod_{j=1}^{g}(\lambda_{2i-1} -\mu_{j}(0)) , \nonumber \\
& &f_{2i-1}^{2}=\prod_{j=1}^{g}(\lambda_{2i-1} - \mu_{j}(x))/
\prod_{j=1}^{g}(\lambda_{2i-1} -\mu_{j}(0)) ,\, i=1,\ldots,g \nonumber
\end{eqnarray}
then we have the following expansion of the potential $u(x)$ in terms of
squares of antiperiodic eigenfunctions
\begin{equation}
u(x)=2\sum_{i=1}^{g}\,f_{2i-1}^2 e_{2i-1}^{-2} +2
\sum_{i=1}^{g}\lambda_{2i-1} -\sum_{i=0}^{2g}\lambda_{i} .
\end{equation}
We call the dynamical systems such in (i) , (ii), complementary dynamical
systems.
\subsection{Solutions in terms of auxiliary spectrum of Hill's equation}
The solutions of the Garnier system in terms of auxiliary spectrum $%
\mu_{j}(x)$, $j=1,\ldots,g$ are
\begin{eqnarray}
&&\xi_{i} =\xi_{i}^{0}[ F(x,a_{i})/F(0,a_{i})]^{1/2}\,
\exp\left(\int_{0}^{x}\,\sqrt{R(a_{i})}/F(x^{\prime},a_{i})dx^{\prime}%
\right) , \label{soluG1} \\
&&\eta_{i}=\eta_{i}^{0}[ F(x,a_{i})/F(0,a_{i}]^{1/2}\,
\exp\left(-\int_{0}^{x}\,\sqrt{R(a_{i})}/F(x^{\prime},a_{i})dx^{\prime}%
\right) , \label{soluG2}
\end{eqnarray}
where $\mu_{j}(x)$ satisfies the following system of differential equations
\begin{equation}
\frac{d}{dx}\mu_{j}(x) = 2\sqrt{R(\mu_{j})}/\prod_{j\neq k}^{g}(\mu_{j}(x) -
\mu_{k}(x)) , \label{muSys}
\end{equation}
with initial conditions
\begin{equation}
\mu_{j}(0)\in [\lambda_{2i-1},\lambda_{2i}],\quad
\xi_{i}^{0}\eta_{i}^{0}=F(0,a_{i})/\prod_{i\neq j}^{g}(a_{i}-a_{j}).
\end{equation}
Differentiating expressions
\begin{eqnarray}
\xi_{i}\eta_{i}=\xi_{i}^{0}\eta_{i}^{0}\Psi(x,p_{i})\Psi^{\tau}(x,p_{i}) =
\prod_{j=1}^{g}(\lambda(p_{i})-\mu_{j}(x))/\tilde{A}^{\prime}(p_{i}) ,
\end{eqnarray}
and
\begin{equation}
\xi_{ix}\eta_{i}-\xi_{i}\eta_{ix} =
\left\{\Psi(x,p_{i}),\Psi^{\tau}(x,p_{i})\right\}\xi_{i}^{0}\eta_{i}^{0} ,
\end{equation}
we have
\begin{eqnarray}
\Upsilon(x,P)|_{\lambda=\lambda(p_{i})} = \frac{d}{dx} \log \xi_{i}(t)
\nonumber \\
=[\frac{1}{2}\frac{d}{dx}\,\prod_{j=1}^{g}\,(\lambda -\mu_{j}(x)) +\sqrt{%
R(\lambda)}]/\prod_{j=1}^{g}\,(\lambda -\mu_{j}(x))
|_{\lambda=\lambda(p_{i})} \label{chi1}
\end{eqnarray}
\begin{eqnarray}
\Upsilon^{\tau}(x,P)|_{\lambda=\lambda(p_{i})} = \frac{d}{dx} \log
\eta_{i}(t) \nonumber \\
=[\frac{1}{2}\frac{d}{dx}\,\prod_{j=1}^{g}\,(\lambda -\mu_{j}(x)) -\sqrt{%
R(\lambda)}]/\prod_{j=1}^{g}\,(\lambda -\mu_{j}(x))
|_{\lambda=\lambda(p_{i})} \label{chi2}
\end{eqnarray}
direct integration of (\ref{chi1}), (\ref{chi2}) gives the solutions (\ref
{soluG1}), (\ref{soluG2}). The function $\Upsilon(x,P)$ has $g$ poles at $%
\mu_{j}(x)$, then the numerator of (\ref{chi1}) is zero when $%
\lambda=\mu_{j}(x)$ and the following system takes place
\begin{equation}
\frac{d}{dx}\left(\prod_{j=1}^{g}\,(\lambda -\mu_{j}(x))\right)
|_{\lambda=\mu_{j}(x)} = 2\sqrt{R( \mu_{j}(x)) } .
\end{equation}
This is another form of the system (\ref{muSys}). In the same way, we can
obtain the solutions of the coupled Neumann system by replacing $%
\lambda(p_{i})$, $i=1,\ldots,g$ with $\lambda(q_{i})$, $i=0,\ldots,g$ in (%
\ref{soluG1}), (\ref{soluG2}).
Let $\lambda(p_{i})$ be the antiperiodic eigenvalues $\lambda_{2i-1}$, $%
i=1,\ldots,g$. Then the exponential function in (\ref{soluG1}), (\ref{soluG2}%
) cancel, $\xi_{i}^{0}=\eta_{i}^{0}$ and the solutions of the $g$%
-dimensional oscillator are
\begin{equation}
\xi_{i}^{2}=\prod_{j=1}^{g}\,(\lambda_{2i-1} - \mu_{j}(x))/ \prod_{i\neq
j}^{g}\,(\lambda_{2i-1}-\lambda_{2j-1}) .
\end{equation}
Let $\lambda(p_{i})=a_{i}$ be in a general position, i.e. $\lambda(p_{i})\in
[\lambda_{2i-1},\lambda_{2i}] $ and by Deift elimination procedure we may
identify $\xi_{i}$ with $\xi_{i}^{0}[F(x,a_{i})/F(0,a_{i})]^{1/2}$ and
\begin{equation}
\theta_{i}=\int_{0}^{x}\,\sqrt{R(a_{i})}/\prod_{i=1}^{g}(a_{i}-\mu_{j}(x^{%
\prime}))dx^{\prime}, \quad \xi_{i}^{0}=\eta_{i}^{0} ,
\end{equation}
and, hence, the solutions of the Rosochatius-Wojciechowski system are
\begin{equation}
\xi_{i}^2=\prod_{j=1}^{g}\,(a_{i}-\mu_{j}(x))/ \prod_{i\neq
j}^{g}(a_{i}-a_{j}) . \label{GarnT}
\end{equation}
Inserting explicit expression of BA-function given by (\ref{BA1}) in Hill's
equation we have
\begin{eqnarray}
\frac{1}{2} F_{xx}(x,\lambda)F(x,\lambda) - \frac{1}{4} F_{x}^{2}(x,%
\lambda)-(u(x)+\lambda)F^{2}(x,\lambda)= -R(\lambda) . \label{qEQ}
\end{eqnarray}
The polynomial solution of (\ref{qEQ}) below we will call Novikov polynomial
\cite{Nov}. Assuming that Novikov polynomial dependts on time $t$,
the zero curvature representation for KdV hierarchy of equations
have the following form
\begin{equation}
M_{t}(\lambda')-L_x(\lambda^{\prime})+\left[M(\lambda`),L(\lambda^{\prime})
\right ] , \label{ZCeq}
\end{equation}
where matrices $L$ and $M$ are given by
\begin{eqnarray}
L(\lambda^{\prime})&=&\left(
\begin{array}{cc}
-F_{x}(x,t,\lambda^{\prime})/2 & F(x,t,\lambda^{\prime}) \\
-F_{xx}(x,t,\lambda^{\prime})/2 +Q(x,t,\lambda')F(x,t,\lambda') &
F_{x}(x,t,\lambda^{\prime})/2
\end{array}
\right) \nonumber \\
M^{\prime}(\lambda^{\prime})&=&\left(
\begin{array}{cc}
0 & 1 \\
Q(x,t,\lambda') & 0
\end{array}
\right)\,. \nonumber
\end{eqnarray}
The equation (\ref{ZCeq}) is equivalent to
\begin{equation}
\frac{\partial Q}{\partial t}= -2\left[\frac{1}{4}\partial^3_x
-Q(x,t,\lambda^{\prime})\partial_x -\frac{1}{2} Q_{x}(x,t,\lambda^{\prime})
\,\right]\cdot F(x,\lambda^{\prime})\,. \label{Geq}
\end{equation}
where $Q(x,t,\lambda^{\prime})=u(x)+\lambda^{\prime}$
in the case of KdV hierarchy. Equation (\ref{Geq}) is called the generating
equation. For a different choices of the form of $F(x,t,\lambda^{\prime})$ and
$Q(x,t,\lambda^{\prime})$, this procedure leads to different hierarchies of
integrable equations, as an example to the KdV, nonlinear Shr\"{o}dinger and
sine-Gordon hierarchies or to the Dym hierarchy \cite{ACHM}.
The Lax representation $L_{x}=[M,L]$ yields the hyperelliptic curve
$K=(\mu^{\prime},\lambda^{\prime})$
\begin{eqnarray}
&&\mbox{Det}(L(\lambda^{\prime})-\mu^{\prime}\mbox{I})=0 , \\
&&\mu^{2}=-\frac{1}{2}F F_{xx}+\frac{1}{4}F_{x}^{2}+(\lambda'+u)F=R(\lambda') ,
\nonumber
\end{eqnarray}
generating the integrals of motion for stationary KdV hierarchy.
Using the equation (\ref{qEQ}) and the following expansion of potential
$u(x)$
in terms of squares of eigenvalue functions $\xi_{k}^{2}(x)=\beta_{k}(x)$
\begin{equation}
u(x)=2\sum_{i=1}^{g}\xi_{k}^2(x)+2\sum_{i=1}^{g} a_{i}
-\sum_{k=0}^{2g}\lambda_{i} ,
\end{equation}
have the form
\begin{eqnarray}
\frac{1}{2}\beta_{kxx}\beta_{k}-\frac{1}{4}\beta_{k}^{2}-
(u(x)+a_{k})\beta_{k}^{2}=-\frac{R(a_{k})}{a_{kj}}=-f_{k} \label{SolGarnT}
\end{eqnarray}
where we use the solutions $F(x,a_{k})/a_{kj}$ of Rosochatius-Wojciechowski
system and $a_{kj}\equiv\prod_{k\neq j}^{g}(a_{k}-a_{j})$. Denoting $%
f_{k}=R(a_{k})/a_{kj}$ and $d=\sum_{i=1}^{g}a_{i}-\frac{1}{2}%
\sum_{k=0}^{2g}\lambda_{k}$ for the original variable $\beta_{k}=\xi_{k}^{2}$
we have the following equation
\begin{eqnarray}
\xi_{kxx}=2\left(\sum_{i=1}^{g}\xi_{k}^{2} +d\right)\xi_{k}-\frac{f_{k}}{%
\xi_{k}^{3}} .
\end{eqnarray}
To understand the role of GD polynomials and of their generating function
in the construction of a map between stationary and restricted
flows of KdV equation and exact solution of completely integrable systems
related to Hill's equation let us consider the following system:
\begin{eqnarray}
&&p_{00}-v_{1}=a_{0}, \quad P_{0}\left(v_{1}-\sum_{j=1}^{n}\beta_{j}\right),
\\
&&P^{\lambda_{k}}\beta_{k}=0, \qquad (k=1,\ldots,n)
\end{eqnarray}
where$\lambda_{1},\ldots,\lambda_{n}$ are fixed parameters, $%
P^{\lambda_{k}}:=P_{1}+4\lambda_{k} P_{0}$ ($P_{0}$ and $P_{1}$ being the
two KdV Poisson operators). This is a system of $(g+2)$ equations in $u,
v_{1},\beta_{1},\ldots,\beta_{g}$. The second equation will be referred to
as the $P_{0}$-restriction of the first KdV flow $X_{0}=P_{0}v_{1}=v_{1x}$,
and the last $n$ equations define the kernel of $g$ Poisson operators
extracted from the Poisson operator. On account of (\ref{GD}), (\ref{Opers})
and (\ref{Brel}) this system is equivalent to the following one:
\begin{eqnarray}
u=\frac{v_{1}}{2}+\frac{a_{0}}{4},\quad v_{1}=\sum_{j=1}^{n}\beta_{j} +
c,\quad B^{\lambda_{k}}(\beta_{k},\beta_{k})=f_{k} ,
\end{eqnarray}
where $c$ and $f_{k}$ are free parameters and $B^{\lambda}$ is just the
generating function of the GD polynomials.
Using the first two equations to eliminate $u$ and $v_{1}$ from the last $g$
equations, one gets a system of $n$ second-order ODEs for $%
\beta_{1},\ldots,\beta_{g}$ :
\begin{eqnarray}
2\beta_{kxx}\beta_{k}-\beta_{kx}^{2}+2\beta_{k}^{2}
\left(\sum_{j=1}^{n}\beta_{j} +d\right) -\lambda_{k}\beta_{k}^{2} = f_{k},
\quad k=1,\ldots,g \label{QRos}
\end{eqnarray}
where $d:=c+a_{0}/2$. Introducing the so-called eigenfunction variables $%
\psi_{j}^{2}=\beta_{j}$ and the momenta $\chi_{j}=\psi_{jx}$, equations (\ref
{QRos}) can be written in canonical Hamiltonian form
\begin{eqnarray}
\psi_{jx}=\frac{\partial K_{G}}{\partial \chi_{j}}, \quad \chi_{jx}=-\frac{%
\partial K_{G}}{\partial \psi_{j}}, \quad j=1,\ldots,n
\end{eqnarray}
with Hamiltonian
\begin{eqnarray}
K_{G}=\sum_{j=1}^{n}\chi_{j}^{2}
-\left(\sum_{k=1}^{n}\psi_{j}^{2}\right)^{2}- \sum_{j=1}^{n}
a_{j}\psi_{j}^{2}- \sum_{j=1}^{n}\frac{f_{j}}{\psi_{j}^{2}} .
\end{eqnarray}
A set of integrals of motion is
\begin{eqnarray}
I_{j}=&&\chi_{j}^{2}-\psi_{j}^{2}\left(a_{j}+
\sum_{k=1}^{g}\psi_{k}^{2}\right)- \frac{f_{j}}{\psi_{j}^{2}}+ \nonumber \\
&& \sum_{k\neq j}^{n}\frac{1}{a_{jk}} \left(-\frac{f_{j}\psi_{k}^{2}}{%
\psi_{j}^{2}} - \frac{f_{k}\psi_{j}^{2}}{\psi_{k}^{2}}+
(\psi_{j}\chi_{k}-\psi_{k}\chi_{j})^{2}\right) . \label{IntRosI}
\end{eqnarray}
where we denote $a_{jk}=a_{j}-a_{k}$.
Now we shall construct a map between the $g$-th
stationary flow and the previous restricted flow of the KdV hierarchy. To
this end we extend the corresponding phase spaces, regarding some free
parameters in the Hamiltonian functions as addiitional dynamical variables.
As for the $P_{1}$-formulation of the stationary flow we extend its phase
space to a $(3g+1)$-dimensional space, $\tilde{M}_{n}$, with coordinates $%
(q_{k},p_{k};a_{0},\ldots,a_{g-1},a_{2g})$; analogously we consider the $%
P_{0}$-formulation of the first restricted flow in the extended space $%
\tilde{{\it M}}_{g}$ with coordinates $(\psi_{k},\chi_{k};f_{1},%
\ldots,f_{k},d)$.
Let us consider the solutions $q_{k}$ of the dynamical equations (\ref{Deq}%
); then $v^{(g)}(\lambda)$ given by
\begin{equation}
v^{(g)}(\lambda)=\lambda(q^{2}(\lambda))^{(g-1)}-q_{g}^{2} ,
\end{equation}
with $q(\lambda)=1+\sum_{j=1}^{g}\,q_{j}\lambda^{-j}$, satisfies (2.17), and
consequently satisfies the following equation:
\begin{equation}
B^{\lambda}(v^{(g)},v^{(n)})=\lambda^{2g} \mbox{d}(\lambda) ,
\end{equation}
where, as above, we put $u=v_{1}/2+a_{0}/4$. So, for each $g$-tuple of
distinct complex parameters $a_{j}$, any solution $v^{(g)}(\lambda)$
fulfills the system
\begin{equation}
B^{a_{k}}(v^{(g)}(a_{k}),v^{(g)}(a_{k})=a_{k}^{2g} %
\mbox{d}(a_{k}) , \quad (k=1,\ldots,g)
\end{equation}
where $v^{(g)}(a_{k}):=v^{(\lambda)}|_{\lambda=a_{k}}$. In order to have
a solution also satisfying the second equation $v_{1}=
\sum_{j=1}^{g}\beta_{j} +c$, the Lagrange interpolation formula can be used.
It allows us to represent the polynomial $v^{n}(\lambda)$ by
\begin{equation}
v^{(n)}(\lambda)=a(\lambda)\left(1+\sum_{j=1}^{g}\frac{\beta_{j} }
{\lambda-a_{j} }\right) , \label{LagIF}
\end{equation}
where $a(\lambda)=\prod_{j=1}^{g}(\lambda-a_{j})$, and
\begin{equation}
\beta_{j}=\frac{v^{(g)}(a_{k})}{a^{\prime}(a_{k})},
(k=1,\ldots,g) .
\end{equation}
($a^{\prime}(\lambda)$ means the derivative of $a(\lambda)$ with respect to $%
\lambda$.
Obviously the $g$ functions $\beta_{k}$ are solutions of the following
system
\begin{eqnarray}
2\beta_{kxx}\beta_{k}-\beta_{kx}^{2}+2\beta_{k}^{2}
\left(\sum_{j=1}^{g}\beta_{j} +d\right) -\lambda_{k}\beta_{k}^{2} = \frac{%
\lambda_{k}^{2g}\mbox{d}(a_{k})}{(a^{\prime}(a_{k}))^{2}} , \quad
k=1,\ldots,g \label{QRos1}
\end{eqnarray}
Furthermore, $\beta_{k}$ satisfy the so-called Bargmann constraint
\begin{equation}
\sum_{j=1}^{g} (\beta_{j}-a_{j}) = v_{1} ,
\end{equation}
as one can verify by means of (\ref{LagIF}).
The function $B^{\lambda}$ is also a generating function of integrals of
motion for Garnier system. Indeed evaluating the function $B^{\lambda}$ by
means of (\ref{LagIF}) and eliminating the first $x$-derivatives of $%
\chi_{k} $ by means of Hamilton equations (\ref{Deq}), one gets
\begin{eqnarray}
4\sum_{j=1}^{g}\frac{I_{j}}{\lambda-\lambda_{j}} + \sum_{j=1}^{g} \frac{f_{j}%
}{(\lambda-\lambda_{j})^{2}} + 2d-\lambda=\frac{\lambda^{2g}\hat{a}(\lambda)%
}{(a(\lambda))^{2}} ,
\end{eqnarray}
where $I_{j}$ are the functions. Taking in this equation the residues at $%
\lambda=a_{j}$ it follows that the functions $I_{j}$ are integrals of motion
along the flow (\ref{Deq}).
Let $\lambda(q_{i})=b_{i}$ be in a general position, i.e.
$\lambda(q_{i})=b_{i}\in (-\infty,\lambda_{0}],
[\lambda_{2i-1},\lambda_{2i}] $ and by Deift elimination procedure we may
identify $\tilde{\psi}_{i}$ with $\tilde{\xi}%
_{i}^{0}[F(t,b_{i})/F(0,b_{i})]^{1/2}$ and
\begin{equation}
\tilde{\theta}_{i}=\int_{0}^{x}\,\sqrt{R(b_{i})}/\prod_{i=1}^{g}
(b_{i}-\mu_{j}(x^{\prime}))dx^{\prime},
\end{equation}
and, hence, the solutions of the Rosochatius system are
\begin{equation}
\tilde{\psi}_{i}^2=\prod_{j=0}^{g}\,(b_{i}-\mu_{j}(x))/ \prod_{i\neq
j}^{g}(b_{i}-b_{j}) .
\end{equation}
where $i,j=0,\ldots g$. Now we illustrate the general aproach with some
simple examples
{\bf Example 1} Let $u(x)=6\wp(x+\omega^{\prime})$ be the two-gap Lam\'e
potential with simple periodic spectrum (see for example \cite{EK})
\begin{equation}
\lambda_{0}=-\sqrt{3g_{2}},\quad \lambda_{1}=-3e_{0}\quad ,
\lambda_{2}=-3e_{1}\quad , \lambda_{3}=-3e_{2}\quad , \lambda_{4}=\sqrt{%
3g_{2}}.
\end{equation}
and the corresponding Hermite polynomial have the form
\begin{equation}
F(\wp(x+\omega^{\prime}),\lambda)=\lambda^{2}-
3\wp(x+\omega^{\prime})\lambda+
9\wp^{2}(x+\omega^{\prime})-\frac{9}{4}g_{2} . \label{HerPol}
\end{equation}
Consider the following genus $2$ nonlinear anisotropic oscillator with
Hamiltonian
\begin{equation}
H=\frac{1}{2}(p_{1}^{2}+p_{2}^{2})+\frac{1}{4}(q_{1}^{2}+q_{2}^{2})^{2}-
\frac{1}{2}(a_{1}q_{1}^{2}+a_{2}q_{2}^{2}),
\end{equation}
where $(q_{i},p_{i})$, $i=1,2$ are canonical variables with $p_{i}=q_{ix}$
and $a_{1},a_{2}$ are arbitrary constants. The
simple solutions of these system are given in terms of Hermite polynomial
\begin{equation}
q_1^2=2\frac{F(x,\tilde{\lambda}_{1})} {\tilde{\lambda}_{2}-\tilde{\lambda}%
_{1}} ,\quad q_2^2=2\frac{F(x,\tilde{\lambda}_{2})} {\tilde{\lambda}_{1}-%
\tilde{\lambda}_{2}} ,
\end{equation}
Let us list the corresponding solutions
\begin{itemize}
\item Periodic solutions expressed in terms of single Jacobian elliptic
functions
\end{itemize}
The nonlinear anisotropic oscillator admit the following solutions:
\begin{eqnarray}
q_1 & = & C_1 \mbox{sn}(\alpha x, k), \\
q_2 & = & C_2 \mbox{cn}(\alpha x, k), \label{onegap}
\end{eqnarray}
where amplitudes $C_1$ ,$C_2$ and temporal pulsewidth $1/\alpha$ of are
defined by parameters $a_{1}$ and $a_{2}$ as
\begin{eqnarray}
\alpha^2 k^2 & = & a_{2}-a_{1} \nonumber \\
C^{2}_{1} & = & a_{2} + \alpha^2 - \alpha^2 k^2 , \nonumber \\
C^{2}_{2} & = & a_{1} + \alpha^{2} , \nonumber
\end{eqnarray}
where $0 < k < 1$.
Following our spectral method it is clear, that the solutions \ref{onegap}
are associated with eigenvalues $\lambda_2 = - e_2$ and $\lambda_3 = - e_3$
of one -- gap Lam\'e potential.
\begin{itemize}
\item Periodic solutions expressed in terms of products of Jacobian
elliptic functions
\end{itemize}
\begin{eqnarray}
q_1 & = & C \mbox{dn}(\alpha x, k) \mbox{sn}(\alpha x, k), \\
q_2 & = & C \mbox{dn}(\alpha x ,k ) \mbox{cn}(\alpha x, k), \label{Flor}
\end{eqnarray}
where $\mbox{sn}$,$\mbox{cn}$, $\mbox{dn}$ are the standard Jacobian
elliptic functions \cite{BE}, $k$ is the modulus of the elliptic functions $%
0 < k < 1$, an the wave characteristic parameters: amplitude $C$, temporal
pulsewidth $1/\alpha$ and $k$ are related to the physical parameters and, $k$
through the following dispersion relations
\begin{eqnarray}
C^{2} & = & \frac{2 (4a_{2} - a_{1})}{5} , \nonumber \\
k^{2} & = & \frac{(4a_{2}-a_{1})} {15} , \nonumber \\
\alpha^{2} & = & \frac {5 (a_{2}-a_{1})} {4a_{2}-a_{1}} . \nonumber
\end{eqnarray}
We have found the following solutions of the nonlinear oscillator
\begin{eqnarray}
q_1 & = & C\alpha^2 k^2 \mbox{cn}(\alpha x,k)\mbox{sn}(\alpha x,k) \\
q_2 & = & C\alpha^2 \mbox{dn}^2 (\alpha x, k) + C_{1} \label{UzKos}
\end{eqnarray}
where $C$, $C_1$, $\alpha$ and $k$ are expressed through parameters
$a_{1}$ and $a_{2}$ by the following relations
\begin{eqnarray}
C^2 & = & \frac {18} {a_{2}-a_{1}} , \nonumber \\
C_1 & = & \frac { C (4a_{1}-a_{2})}{5},\quad
\nonumber \\
k^2 & = & \frac {2 \sqrt { \frac{5}{3} (a_{2}^{2}-a_{1}^{2})}}
{2 \sqrt{\frac{5}{3} (a_{2}^{2}-a_{1}^{2})}+aa_{2}-3a_{1}},
\nonumber \\
\alpha^2 & = & \frac {1}{10} ( 2 a_{2}-3a_{1}+
\sqrt{\frac{5}{3}(a_{2}^{2}-a_{1}^2) } ).
\end{eqnarray}
\begin{itemize}
\item Periodic solutions associated with the two-gap Treibich-Verdier
potentials
\end{itemize}
Below we construct the two periodic solutions associated with the
Treibich-Verdier potential. Let us consider the potential
\begin{equation}
u(x)=6\wp(x+\omega^{\prime})+2{\frac{(e_1-e_2)(e_1-e_3)}{\wp(x+\omega^{%
\prime})-e_1}} \label{tv4}
\end{equation}
and construct the solution in terms of Lam\'e polynomials associated with
the eigenvalues $\tilde{\lambda}_1,\tilde{\lambda}_2$, $\tilde{\lambda}_1 >
\tilde{\lambda}_2$
\begin{eqnarray}
\tilde{\lambda}_1&=&e_2+2e_1+2\sqrt{(e_1-e_2)(7e_1+2e_2)}, \nonumber \\
\tilde{\lambda}_2&=&e_3+2e_1+2\sqrt{(e_1-e_3)(7e_1+2e_3)}. \label{zz}
\end{eqnarray}
The finite and real solutions $q_1,q_2$ have the form
\begin{eqnarray}
q_1&=& C_{1}\mbox{sn}(z,k)\mbox{dn}(z,k) +C_{2}\mbox{sd}(z,k) , \nonumber \\
q_2&=& C_{3}\mbox{cn}(z,k)\mbox{dn}(z,k) +C_{4}\mbox{cd}(z,k), \label{ee2}
\end{eqnarray}
where $C_{i}$, $i=1,\ldots 4$ are constants and have important geometrical
interpretation \cite{EK}. The concrete expressions in terms of $k,\tilde{%
\lambda}_{1},\tilde{\lambda_{2}}$ are given in \cite{CEEK}
Analogously we can find the elliptic solution associated with the
eigenvalues
\begin{eqnarray}
\tilde{\lambda}_1&=&e_2+2e_1+2\sqrt{(e_1-e_2)(7e_1+2e_2)},\quad \tilde{%
\lambda}_2=-6e_1, \label{zz2}
\end{eqnarray}
We have
\begin{eqnarray}
q_1&=&\tilde{C}_{1}\mbox{dn}^{2}(z,k) \\
q_2&=& C_{1}\mbox{sn}(z,k)\mbox{dn}(z,k) +C_{2}\mbox{sd}(z,k) , \label{ee3}
\end{eqnarray}
where $C$ is given in \cite{CEEK}.
The general formula for elliptic solutions of genus $2$ nonlinear
anisotropic oscillator is given in \cite{CEEK}
\begin{eqnarray}
q_1^2&=&{\frac{1}{\tilde{\lambda}_2-\tilde{\lambda}_1}}
\left(2\tilde{\lambda}_1^2+2\tilde{\lambda}_1\sum_{i=1}^N
\wp(x-x_i) \right. \nonumber
\\ &&+\left.6\sum_{1\leq i< j\leq N}\wp(x-x_i)\wp(x-x_j)-{\frac{Ng_2}{4}}+
\sum_{1\leq i< j\leq 5}\lambda_i\lambda_j\right), \label{q1} \\
q_2^2&=&{\frac{1}{\tilde{\lambda}_1-\tilde{\lambda}_2}}
\left(2\tilde{\lambda}_2^2+2\tilde{\lambda}_2\sum_{i=1}^N
\wp(x-x_i) \right. \nonumber
\\ &&+\left.6\sum_{1\leq i< j\leq N}\wp(x-x_i)\wp(x-x_j)-{\frac{Ng_2}{4}}+
\sum_{1\leq i< j\leq 5}\lambda_i\lambda_j\right), \label{q2}
\end{eqnarray}
where $x_{i}$ are solutions of equations $\sum_{i\neq j}\wp'(x_{i}-x_{j})=0,
j=1,\ldots, N$ and $N$ is positive integer.
{\bf Example 2} Garnier type system.
Consider the following genus $2$ Garnier type system with
Hamiltonian
\begin{equation}
H=\frac{1}{2}(p_{1}^{2}+p_{2}^{2})+\frac{1}{4}(q_{1}^{2}+q_{2}^{2})^{2}-
\frac{1}{2}(\gamma_{1} q_{1}^{2}+\gamma_{2} q_{2}^{2})+\frac{f_{1}}{q_{1}^{2}}+
\frac{f_{2}}{q_{2}^{2}},
\end{equation}
where $(q_{i},p_{i})$, $i=1,2$ are canonical variables with $p_{i}=q_{ix}$
and $a_{1},a_{2}$ are arbitrary constants. The
simple solutions of these system are given in terms of Hermite polynomial
in the following form (\ref{GarnT})
\begin{equation}
q_1^2=2\frac{F(x,a_{1})}
{a_{1}-a_{2}} , \quad
q_2^2=2\frac{F(x,a_{2})}
{a_{2}-a_{1}} ,
\end{equation}
the same settings of periodic spectra, Lam\'e potential and Hermite
polynomials (\ref{HerPol})
as in example 2. The main results from the general theory are the following:
i) the parameters $f_{1}$ and $f_{2}$ are expressed in terms of algebraic
curve by (\ref{SolGarnT})
\begin{eqnarray}
f_{1}=\frac{R(a_{1})}{a_{1}-a_{2}}, \quad
f_{2}=\frac{R(a_{2})}{a_{2}-a_{1}}, \nonumber \\
R(\lambda)=(\lambda^{2}-3g_{2})(\lambda+3e_{1})(\lambda+3e_{2})(\lambda+3e_{3})
. \nonumber
\end{eqnarray}
ii) the parameters
$a_{1}$ and $a_{2}$ must lie in one or other of the intervals
\begin{equation}
[-\sqrt{3g_{2}}, 3e_{1}], [3e_{2},3e_{3}], [\sqrt{3g_{2}},\infty)
\end{equation}
and $\gamma_{1}=3a_{1}+2a_{2}, \gamma_{2}=2a_{1}+3a_{2}$.
These results are in complete agreement with solutions obtained by different
method in recent paper \cite{pp98}.
{\bf Example 3} Simple solutions of the H\'enon-Heiles type system.
We consider a generalized H\'enon-Heiles type system with two-degrees of
freedom.
Its Hamiltonian is
\begin{equation}
H_{0}=\frac{1}{2}(p_{1}^{2}+p_{2}^{2}) + q_{1}^{3}+\frac{1}{2}q_{1}
q_{2}^{2} + \frac{a_{4}}{8 q_{2}^{2}} +\frac{a_{0}}{2}\left(q_{1}^{2} +\frac{%
1}{4} q_{2}^{2}\right) -\frac{a_{1}}{4} q_{1} , \label{HH}
\end{equation}
where $q_{1}, q_{2}, p_{1}, p_{2}$ are the canonical coordinates and momenta
and $a_{0}, a_{1}, a_{4}$ are free constant parameters. This Hamiltonian
encompasses the two cases $a_{0}=a_{4}=0$ and $a_{0}=a_{1}=0$ introduced in
\cite{18}. Moreover $H_{0}$ is related with the Hamiltonian
\begin{equation}
H_{H}= \frac{1}{2}(p_{1}^{2}+p_{2}^{2}) + \bar{q}_{1}^{3}+\frac{1}{2}\bar{q}%
_{1} \bar{q}_{2}^{2} + \frac{\bar{a_{4}}}{{8 \bar{q}_{2}^{2}}}
+\frac{1}{2}\left(A\bar{q}_{1}^{2} + B \bar{q}_{2}^{2}\right) ,
\end{equation}
through the map
\begin{equation}
q_{1}=\bar{q}_{1} + \frac{A}{2} -2B,\quad q_{2}=\bar{q}_{2}, \quad a_{0}=-2
A+12 B, \quad a_{1}=-A^{2}+16 AB-48 B^{2} .
\end{equation}
The function $H_{H}$ is the Hamiltonian of a classical integrable
H\'enon-Heiles system with the additional term $a_{4}/8\bar{q}%
_{2}^{2}$.
The function (\ref{HH}) is the Hamiltonian of the vector field obtained
reducing $X_{0}(u)=u_{x}$ to the stationary manifold $M_{2}$ given by the
fixed points of the flow $X_{2}+c_{1}X_{1}+c_{2}X_{0}$
\begin{equation}
M_{2}={u|u^{(5)}+10u_{xxx}u+20u_{xx}u_{x}+30u_{x}u^{2}+
c_{1}(u_{xxx}+6u_{x}u) +c_{2}u_{x}=0}
\end{equation}
where $c_{1}=-a_{0}/2$, $c_{2}=-a_{1}/2+a_{0}^{2}/4$.
It can be obtained specializing to the case $g=2$ the Hamiltonian of the $%
P_{1}$-system. In this case $H_{2}^{(1)}=H_{0}$ and the canonical
coordinates and momenta are, respectively, $q_{1}=v_{1}/2$, $%
q_{2}^{2}=-v_{2} $, $p_{1}=q_{1x}$, $p_{2}=q_{2x}$. The integrals of motion
obtained by the reduction of the GD polynomials are
\begin{eqnarray}
&&H_{0}\equiv -\frac{1}{8}p_{02}|_{x} \\
&&H_{2}\equiv -\frac{1}{8}p_{22}|_{x} =-\frac{a_{4}}{8} \\
&&H_{1}\equiv -\frac{1}{8}p_{12}|_{x}=
\end{eqnarray}
where $-\frac{1}{8}p_{12}|_{x}$ is given by
\begin{eqnarray}
p_{2}q_{1}-p_{1}p_{2}q_{2}-\frac{1}{2}q_{1}^{2}q_{2}^{2}-\frac{1}{8}%
q_{2}^{4}+ \frac{a_{4}q_{1}}{4q_{2}^{2}}-\frac{a_{0}}{4}q_{1}q_{2}^{2}+
\frac{a_{1}}{8}q_{2}^{2} .
\end{eqnarray}
Next we will derive $(2\times 2)$ matrix Lax representation for generalized
H\'enon-Heiles system (\ref{HH}). Using Lax representation $L_{x}=[M,L]$,
particular case of eq. (\ref{ZCeq}) i.e. when there is no time dependence, we
have
\begin{eqnarray}
&&F(x,\lambda)=\lambda^2+\frac{1}{2} q_{1}\lambda-\frac{1}{16}q_{2}^{2}, \quad
\quad
V=-F_{x}/2=-\frac{1}{4}p_{1}\lambda + \frac{1}{16} q_{2} p_{2} , \nonumber \\
&&W=-F_{xx}/2+Q F=\lambda^{3}-(\frac{1}{2}q_{1}+\frac{1}{4} a_{0})\lambda^2+
\nonumber \\
&&(\frac{1}{4} q_{1}^{2}+\frac{1}{16} q_{2}^{2}-\frac{1}{16}a_{1}+
\frac{1}{8} a_{0} q_{1})\lambda + \frac{1}{16}p_{2}^{2} +
\frac{1}{64} \frac{a_{4}}{q_{2}^{2}} ,
\quad Q(x,\lambda)=\lambda-q_{1}-\frac{1}{4}a_{0} . \nonumber
\end{eqnarray}
The corresponding algebraic curve have the form
\begin{eqnarray}
\mu^2=\lambda^{5}-\frac{1}{4}a_{0}\lambda^{4}-\frac{1}{16}a_{1}\lambda^{3}
+8 H_{0}\lambda^{2}+32\,H_{1}\lambda-\frac{1}{1024}a_{4} .
\end{eqnarray}
Using explicit expression for Hermite polinomial (\ref{HerPol}) we obtain
the following simple solutions for the system (\ref{HH}):
\begin{eqnarray}
q_{1}=-6\wp(x+\omega'), \qquad
q_{2}^2=-16\,(9\wp(x+\omega')^{2} -\frac{9}{4} g_{2}) .
\end{eqnarray}
where $a_{0}=0, a_{1}=3.4.7 g_{2}, A_{4}=-3^{4}.4^{4} g_{2} g_{3}$.
\section{$2\times 2$ Lax representation and $r$-matrix approach}
\setcounter{equation}{0} The Lax equation for completely integrable systems
discussed in the previous section
\begin{equation}
L_x(\lambda^{\prime})=\left[M(\lambda^{\prime}),L(\lambda^{\prime})\right]%
\,, \label{Lax}
\end{equation}
with matrices $L$ and $M$ given by
\begin{eqnarray}
L(\lambda^{\prime})&=&\left(
\begin{array}{cc}
V(x,\lambda^{\prime}) & U(x,\lambda^{\prime}) \\
W(x,\lambda^{\prime}) & -V(x,\lambda^{\prime})
\end{array}
\right) \label{Llm1} \\
M(\lambda^{\prime})&=&\left(
\begin{array}{cc}
0 & 1 \\
Q(x,\lambda^{\prime}) & 0
\end{array}
\right)\,. \label{Llm2}
\end{eqnarray}
is equivalent to the Garnier system, where $U(x,\lambda^{\prime}),
V(x,\lambda^{\prime}), W(x,\lambda^{\prime}), Q(x,\lambda^{\prime})$ have
the form
\begin{eqnarray}
U(x,\lambda^{\prime})=a(\lambda^{\prime})\left(1-\sum_{i=1}^{g} \frac{%
\xi_{i}\eta_{i}}{\lambda^{\prime}- a_{i}}\right), \,
V(x,\lambda^{\prime})=-\frac{1}{2}U_{x}(x,\lambda^{\prime}) \\
W(x,\lambda^{\prime})=a(\lambda^{\prime})\left(\lambda^{\prime}+%
\sum_{i=1}^{g}\,\xi_{i}\eta_{i}+ \sum_{i=1}^{g}\frac{\xi_{ix}\eta_{ix}}{%
\lambda^{\prime}- a_{i}}\right), \quad
Q(x,\lambda^{\prime})=\lambda^{\prime}+ 2\sum_{i=1}^{g}\xi_{i}\eta_{i} .
\end{eqnarray}
Finally we point out one usefull expression, which is easy to derive from
Lax representation (\ref{Lax})
\begin{equation}
W(x,\lambda^{\prime})=U(x,\lambda^{\prime})Q(x,\lambda^{\prime})-\frac{1}{2}
U(x,\lambda^{\prime})_{xx}. \label{W}
\end{equation}
The Lax representation yields the hyperelliptic curve $K=(\mu^{\prime},%
\lambda^{\prime})$
\begin{equation}
\mbox{Det}(L(\lambda^{\prime})-\mu^{\prime}\mbox{I})=0 ,
\end{equation}
generating the integrals of motion $H, F^{(i)}, i=1,\ldots,g$. We have
\begin{equation}
\mu^{2}=V^{2}(x,\lambda^{\prime})+U(x,\lambda^{\prime})
W(x,\lambda^{\prime}) , \label{Lcurve}
\end{equation}
From (\ref{Lcurve}) and explicit expressions of $U(x,\lambda^{\prime}),
V(x,\lambda^{\prime}), W(x,\lambda^{\prime})$ we obtain
\begin{equation}
\mu^{2}=a(\lambda^{\prime})^{2}\left(\lambda^{\prime}+ \sum_{i=1}^{g}\frac{%
H_{i}}{\lambda^{\prime}-a_{i}} + \frac{1}{4}\sum_{i=1}^{g}\frac{J_{i}^{2}}{%
(\lambda^{\prime}-a_{i})^{2}} +\frac{1}{2}\sum_{i=1}^{g}\frac{I_{i}}{%
\lambda^{\prime}- a_{i}}\right) ,
\end{equation}
where
\begin{eqnarray}
I_{i}&=&\sum_{k\neq i}\frac{
(\xi_{k}\eta_{ix}-\eta_{i}\xi_{kx}) (\eta_{k}\xi_{ix}-\xi_{i}\eta_{kx})}{%
a_{k}-a_{i}} , \nonumber \\
&+&\sum_{k\neq i}\frac{
(\xi_{i}\xi_{kx}-\xi_{k}\xi_{ix}) (\eta_{i}\eta_{kx}-\eta_{k}\eta_{ix})}{%
a_{k}-a_{i}} , \nonumber \\
&&H_{i}=\xi_{ix}\eta_{ix}-a_{i}\xi_{i}\eta_{i}-\xi_{i}\eta_{i}
\left(\sum_{k=1}^{g}\xi_{k}\eta_{k}\right) , \,
J_{i}=\xi_{ix}\eta_{i}-\xi_{i}\eta_{ix},
\end{eqnarray}
and $\sum_{i=1}^{g}H_{i}$ is the Hamiltonian for Garnier system. Simple
reduction $\eta_{i}= \xi_{i}^{*}$ gives us the second flow of stationary
vector nonlinear Schr\"{o}dinger equation, where by $*$ we denote complex
conjugation. The complementary to the last system is complex Neumann system
(see for example \cite{IA})
\begin{eqnarray}
\tilde{\xi}_{ixx}+2\left(\sum_{j=0}^{g} b_{j}|\tilde{\xi}_{j}|^{2} +|\tilde{%
\xi}_{jx}|^{2}\right)\tilde{\xi}_{i}=b_{i}\tilde{\xi}_{i} .
\end{eqnarray}
with $\sum_{i=0}^{g}|\xi_{i}|^{2}=1$.
Using the Deift elimination procedure we obtain new $2\times 2$ Lax pair for
Rosochatius-Wojciechowski system. Bellow we list only the final results for
Lax pair elements of considered in this paper dynamical systems:
\begin{itemize}
\item Rosochatius-Wojciechowski system
\begin{eqnarray}
&&U(x,\lambda^{\prime})=a(\lambda^{\prime})\left(1-\sum_{i=1}^{g} \frac{%
\psi_{i}^{2}}{\lambda^{\prime}- a_{i}}\right), \,
V(x,\lambda^{\prime})=-\frac{1}{2}U_{x}(x,\lambda^{\prime}) \nonumber \\
&&W(x,\lambda^{\prime})=a(\lambda^{\prime})\left(\lambda^{\prime}+%
\sum_{i=1}^{g}\,\psi_{i}^{2} +
\sum_{i=1}^{g}\frac{1}{%
\lambda^{\prime}- a_{i}}
(\psi_{ix}^{2}-\frac{f_{i}}{\psi_{i}^{2}})\right) ,
\\ &&
Q(x,\lambda^{\prime})=\lambda^{\prime}+ 2\sum_{i=1}^{g}\psi_{i}^{2} .
\nonumber
\end{eqnarray}
\item Rosochatius system
\begin{eqnarray}
&&U(x,\lambda^{\prime})=b(\lambda^{\prime})\left(\sum_{i=1}^{g} \frac{%
\psi_{i}^{2}}{\lambda^{\prime}- b_{i}}\right), \,
V(x,\lambda^{\prime})=-\frac{1}{2}U_{x}(x,\lambda^{\prime}) \nonumber \\
&&W(x,\lambda^{\prime})=b(\lambda^{\prime})\left(1-%
+ \sum_{i=1}^{g}\frac{1}{%
\lambda^{\prime}- b_{i}}
(\psi_{ix}^{2}-\frac{f_{i}}{\psi_{i}^{2}})\right) ,
\\ &&
Q(x,\lambda^{\prime})=\lambda^{\prime}+ 2\sum_{i=1}^{g}\psi_{i}^{2} ,
\nonumber
\end{eqnarray}
where $b(\lambda^{\prime})=\prod_{i=0}^{g}(\lambda^{\prime}-b_{i})$.
\item second stationary flow of vector NLSE
\begin{eqnarray}
&&U(x,\lambda^{\prime})=a(\lambda^{\prime})\left(1-\sum_{i=1}^{g} \frac{%
|\xi_{i}|^{2}}{\lambda^{\prime}- a_{i}}\right), \,
V(x,\lambda^{\prime})=-\frac{1}{2}U_{x}(x,\lambda^{\prime})\nonumber \\
&&W(x,\lambda^{\prime})=a(\lambda^{\prime})\left(\lambda^{\prime}+%
\sum_{i=1}^{g}\,|\xi_{i}|^{2}+ \sum_{i=1}^{g}\frac{|\xi_{ix}|^{2}}{%
\lambda^{\prime}- a_{i}}\right),
\\ &&
Q(x,\lambda^{\prime})=\lambda^{\prime}+ 2\sum_{i=1}^{g}|\xi_{i}|^{2} .
\nonumber
\end{eqnarray}
\item complex Neumann system
\begin{eqnarray}
&&U(x,\lambda^{\prime})=b(\lambda^{\prime})\left(\sum_{i=1}^{g} \frac{%
|\xi_{i}|^{2}}{\lambda^{\prime}- b_{i}}\right), \,
V(x,\lambda^{\prime})=-\frac{1}{2}U_{x}(x,\lambda^{\prime})\nonumber \\
&&W(x,\lambda^{\prime})=a(\lambda^{\prime})\left(1-%
\sum_{i=1}^{g}\frac{|\xi_{ix}|^{2}}{%
\lambda^{\prime}- b_{i}}\right),
\\ &&
Q(x,\lambda^{\prime})=\lambda^{\prime}+ 2\sum_{i=1}^{g}|\xi_{i}|^{2} .
\nonumber
\end{eqnarray}
\end{itemize}
Finally we want to point out that $I_{i}$
for Rosochatius-Wojciechowski system coincide with expression given in
(\ref{IntRosI}). Another Lax equation have the following form
\[
L_x(\lambda^{\prime})=\left[M(\lambda`),L(\lambda^{\prime})\right]\,,
\]
where matrices $L$ and $M$ are given by
\begin{eqnarray}
L(\lambda^{\prime})&=&\left(
\begin{array}{cc}
-F_{x}(x,\lambda^{\prime})/2 & F(x,\lambda^{\prime}) \\
-F_{xx}(x,\lambda^{\prime})/2 & F_{x}(x,\lambda^{\prime})/2
\end{array}
\right) \equiv \nonumber \\
&& \left(
\begin{array}{cc}
V(x,\lambda^{\prime}) & U(x,\lambda^{\prime}) \\
W^{\prime}(x,\lambda^{\prime}) & -V(x,\lambda^{\prime})
\end{array}
\right) \, , \label{KdV1} \\
M^{\prime}(\lambda^{\prime})&=&\left(
\begin{array}{cc}
0 & 1 \\
0 & 0
\end{array}
\right)\,. \label{KdV2}
\end{eqnarray}
where we made the following identification $U(x,\lambda^{\prime})=F(x,%
\lambda^{\prime})$. The Poisson bracket relations for the matrix $%
L(\lambda^{\prime})$ \cite{KRT,T1} are closed into the following $r$-matrix
algebra
\begin{equation}
\left\{{L_{1}}(\lambda^{\prime}),{L_{2}}(\mu^{\prime})\right\}=
[r_{12}(\lambda^{\prime},\mu^{\prime}),{L_{1}}(\lambda^{\prime})]-[r_{21}(%
\lambda^{\prime},\mu^{\prime}), {L_{2}}(\mu^{\prime})\,]\, . \label{Lalg}
\end{equation}
Here the standard notations are introduced:
\begin{eqnarray} \label{r1}
&&L_{1}(\lambda^{\prime})= L(\lambda^{\prime})\otimes I\,,\qquad
L_{2}(\mu^{\prime})=I\otimes L(\mu^{\prime})\,, \nonumber \\
\\
&&r_{12}(\lambda,\mu)=\frac{\Pi}{\lambda-\mu}\,\qquad
r_{21}(\lambda,\mu)=\Pi r_{12}(\mu,\lambda)\Pi\,, \label{r2}
\end{eqnarray}
and $\Pi$ is the permutation operator of auxiliary spaces.
The Poisson bracket relations for the Lax matrix $L(\lambda^{\prime})$ are
preassigned by the initial symplectic structure. It is necessary to
calculate only two brackets
\begin{equation}
\left\{F(x,\lambda^{\prime}),F(x,\mu^{\prime})\right\}=0\,, \label{Br1}
\end{equation}
and
\begin{eqnarray} \label{Br2}
\left\{V(x,\lambda^{\prime}),F(x,\mu^{\prime})\right\}&=& \nonumber \\
&&\left\{ F(x,\lambda^{\prime})\sum_{j=1}^g \frac{\mbox{g}_{jj}\,p_j(x)}
{\lambda^{\prime}-\mu_{j}(x)}, \prod_{j=1}^g(\mu^{\prime}-\mu_{j}(x))\right\}
\nonumber \\
&=& -F(x,\lambda^{\prime})F(x,\mu^{\prime})\sum_{j=1}^n \frac{\mbox{g}_{jj}}{%
(\lambda^{\prime}-\mu_{j}(x))(\mu^{\prime}-\mu_{j}(x))} \nonumber \\
&=&\frac{F(x,\lambda^{\prime})F(x,\mu^{\prime})}{\lambda^{\prime}-\mu^{%
\prime}} \sum_{j=1}^g \left(\frac{\mbox{g}_{jj}}{\lambda^{\prime}-\mu_{j}(x)}%
- \frac{\mbox{g}_{jj}}{\mu^{\prime}-\mu^{\prime}_{j}(x)}\right) \nonumber \\
&=&\frac{1}{\lambda^{\prime}-\mu^{\prime}}\left[F(x,\mu^{\prime})-
F(x,\lambda^{\prime})\right]\,,
\end{eqnarray}
where we used a standard decomposition of rational function
\[
F(x,\lambda^{\prime})^{-1}= \sum_{j=1}^n\frac{\mbox{g}_{jj}}{%
\lambda-\mu_{j}(x)}\,,\qquad \mbox{g}_{jj}=\left.\mbox{Res}%
\right|_{\lambda^{\prime}=\mu_{j}(x)} F(x,\lambda^{\prime})^{-1}(\lambda)\,.
\]
and the following definitions
\begin{eqnarray}
&&\mbox{g}_{jj}=\mbox{Res}_{|\lambda=\mu_{j}(x)}F^{-1}(x,\lambda)= \frac{1}{%
\prod_{k\neq j}(\mu_{j}(x)-\mu_{k}(x))}, \\
&&p_{j}(x)=V(x,\lambda)_{|\lambda=\mu_{j}(x)}=\sqrt{R(\mu_{j}(x))}.
\end{eqnarray}
Another Poisson brackets may be directly derived from these brackets and by
definition of the entries of the Lax matrix $L(\lambda)$ via derivative of
the single function $F(x,\lambda^{\prime})$
\begin{eqnarray} \label{Br3}
\left\{V(x,\lambda^{\prime}),V(x,\mu^{\prime})\right\}&=&0\, \nonumber \\
\left\{W(x,\lambda^{\prime}),F(x,\mu^{\prime})\right\}&=&\frac{d}{dx}
\left\{V(x,\lambda^{\prime}),F(x,\mu^{\prime})\right\}= \nonumber \\
&&\frac{2}{\lambda^{\prime}-\mu^{\prime}}\left[V(x,\lambda^{\prime})-V(x,%
\mu^{\prime}) \right]\, , \nonumber \\
\\
\left\{W(x,\lambda^{\prime}),F(x,\mu^{\prime})\right\}&=&-\frac{1}{2}\frac{%
d^2}{dx^{2}} \left\{V(x,\lambda^{\prime}),F(x,\mu^{\prime})\right\}=
\nonumber \\
&&\frac{1}{\lambda^{\prime}-\mu^{\prime}} \left[W(x,\lambda^{\prime})-W(x,%
\mu^{\prime})\right]\, , \nonumber \\
\left\{W(x,\lambda^{\prime}),W(x,\mu^{\prime})\right\}&=&-\frac{1}{2}\frac{%
d^3}{dx^{3}} \left\{V(x,\lambda^{\prime}),F(x,\mu^{\prime})\right\}=0 \, .
\nonumber
\end{eqnarray}
Applying the following transformation directly to the Lax representation $%
L(\lambda^{\prime})$ we obtain a family of the new Lax pairs \cite{KRT,T1}
\begin{eqnarray} \label{LnLax}
&&L^{\prime}(\lambda^{\prime})=L(\lambda^{\prime})-\sigma_-\cdot
\left[\phi(x,\lambda^{\prime})F(x,\lambda^{\prime})^{-1}\right]_{N}\,,
\qquad \sigma_-=\left(
\begin{array}{cc}
0 & 0 \\
1 & 0
\end{array}
\right)\,, \nonumber \\
\\
&&M^{\prime}(\lambda')=M-\sigma_{-}\cdot\left[\phi(x,\lambda')F(x,\lambda')^{-2
} \right]_{N} = \left( \begin{array}{cc}
0 & 1 \\
Q(x,\lambda') & 0
\end{array}
\right)\,. \nonumber
\end{eqnarray}
Here $\phi(x,\lambda^{\prime})$ is a function on spectral parameter and
$[z]_{N}$ means restriction of $z$ onto the ${\rm ad}^*_R$-invariant
Poisson subspace of the initial $r$-bracket. For the rational $r$-matrix we
can use the linear combinations of the following Taylor projections
\begin{equation}
{[ z ]_{N}}=\left[\sum_{k=-\infty}^{+\infty} z_k\lambda^k\,
\right]_{N}\equiv \sum_{k=0}^{N} z_k\lambda^k\,, \label{Cut}
\end{equation}
or the Laurent projections.
New Lax matrix $L^{\prime}(\lambda^{\prime})$ \cite{KRT,T1} obeys the linear
$r$-bracket, where constant $r_{ij}$-matrices substituted by $%
r_{ij}^{\prime} $-matrices depending on dynamical variables.
\begin{eqnarray}
&&r_{12}(\lambda^{\prime},\mu^{\prime})\rightarrow r^{\prime}_{12}=r_{12}
\nonumber \\
&& -\frac{\left( [\phi(\lambda)
F(x,\lambda^{\prime})^{-2}(\lambda^{\prime})]_{N}- [\phi(\mu)
F(x,\mu')^{-2}]_{N}\,\right)} {(\lambda^{\prime}-
\mu^{\prime})}\cdot\sigma_{-}\otimes\sigma_{-} \,. \label{dpr}
\end{eqnarray}
\section{Conclusions}
In this paper we have given new exact solutions of the physically
significant completely integrable dynamical systems. These solutions can be
interpreted as eigenfunctions of suitable differential operator. New example
of complementary dynamical systems (stationary second flow of the vector
nonlinear Schr\"odinger equation and complex Neumann system) is presented.
\vspace{2cm} {\bf\Large Acknowledgements} \\*\\*I wish to thank to Prof.
V. Gerdjikov, who poined out me the papers \cite{7,8,9}, book \cite{v98} and
for valuable discussions.
|
\section{Introduction}
The Sagittarius dwarf galaxy
discovered \cite{ib94,ib95} as a comoving group
($V_{hel}\approx 140 {\rm kms^{-1}}$) with a velocity
dispersion as small as $2 \rm kms^{-1}$,
is
the nearest external galaxy and will be amenable
to detailed abundance studies with VLT and other 8m class telescopes.
These shall be important for two distinct reasons:
on the one hand, we shall be able to study chemical
evolution in an environment different from our Galaxy; on the other hand,
if Sagittarius shows any chemical signature, this should allow us to
identify Sagittarius debris, currently populating the Galactic Halo.
The colour-magnitude diagram of Sgr
shows a wide Red Giant Branch (RGB), which is
interpreted as evidence of a spread in metallicity of Sgr.
Marconi et al \cite{mara} conclude that the RGB of Sgr lies between
that of 47 Tuc ([Fe/H]$=-0.71$) and M2 ([Fe/H]$=-1.58$).
In the following I shall describe the work done with the ESO NTT
to determine spectroscopic metallicities of giants in Sgr and shall
address the potentiality of VLT on this issue.
\section{NTT Observations}
We selected, from the Marconi et al \cite{mara} sample, stars
on the RGB which ought to display the spread in metallicity
present in Sgr. We used the ESO NTT telescope with the EMMI instrument
in Multi Object Spectrocopy (MOS) mode. The dispersing element was
grism \# 5 providing a resolution of about 1500.
We have obtained spectra for a total of 57 objects in field Sgr1 of
\cite{mara} on June 19th 1996.
In addition, we obtained long slit spectra of one
of the stars observed with the MOS and of star HD 190287
on September 18th 1998.
\begin{figure}
\centering
\includegraphics[]{pl_histo_fe.epsi}
\caption[]{Metallicity distribution for 22 radial velocity
members of Sgr }
\label{eps1}
\end{figure}
Of the 57 stars observed 23 matched the criterion
100 $\rm kms^{-1}\le v_{hel}\le 180 ~ kms^{-1}$ \cite{ib97}, which we adopted,
to ascribe membership to Sagittarius.
For these 23 stars the average radial velocity is
$136\rm ~kms^{-1}\pm 18 ~kms^{-1}$,
in good agreement with previous determinations; the r.m.s.
is dominated by the error in the measure of the radial velocity
and not by the velocity dispersion of Sgr.
\section{Abundance estimates from low resolution spectra}
Our approach has
been to define spectral indices which measure some prominent spectral
features which may be used for abundance estimates.
We defined six spectral indices, two of which measure the Mg I b triplet
and the remaining four measure essentially iron and iron-peak elements.
The location of the indices may be found in Fig.1 of reference\cite{marb} .
The indices are all measured with respect to a common pseudo-continuum,
defined by six quasi-continuum windows.
The aim is to be able to determine
both [Fe/H] and [Mg/Fe].
We computed the values of the indices for a small grid of synthetic
spectra (T=4750 K, 5000 K, 5250 K; log g = 2.50
[Fe/H]$=-1.5,-1.0,-0.5,0.0,+0.5$ ; $\rm \xi=2kms^{-1}$) computed
with the SYNTHE code \cite{Kur93}.
The spectra where broadened with a gaussian profile
of 210 $\rm kms^{-1}$ to match the resolution of
the observed spectra.
The input model-atmospheres were
computed with version 9 of the ATLAS code\cite{Kur93} switching
off the overshooting option. Models with [Fe/H]$=-1.5$ and $-1.0$
were computed using $\alpha-$enhanced opacity distribution functions.
For each star
the temperature was determined from the $(V-I)_0$ colour\cite{mara}
through the calibration of\cite{alo} .
Strictly speaking this calibration is valid only for dwarf stars,
however the $V-I$ colour has only a weak dependence on gravity,
especially in the colour range we are interested in.
We excluded from analysis SgrM 172 ($V-I=1.81$), because it is much
cooler than all the other stars.
The metallicity was
determined from the four iron-sensitive
indices by interpolating in the table of computed indices.
\subsection{Zero-Points}
We made the same analysis for the Sun using the solar atlas\cite{Kur84},
degraded at the same resolution as our
grism spectra, as an observed spectrum and a small
grid of synthetic spectra appropriate for the Sun
(T=5777, log g =4.44, [Fe/H]$=-0.5,0.0,+0.5$).
The difference between these derived metallicities and the
meteoritic
iron abundance, is small
(around 0.1--0.2 dex).
On the assumption
that these differences reflect inadequacies of
our atomic data and model atmosphere and
that the systematics is the same for the Sun and for our program
stars, we treated these as zero-point shifts.
\subsection{Cross-checks}
We determined the metallicity of the Sun from
a twilight spectrum taken with the MOS. The deduced
[Fe/H] was $-0.32,-0.04,-0.01,+0.10$ for the four indices
respectively. These
are not 0.00 because of the noise in the
observed spectrum. One
of the indices (F1) yields an abundance considerably
different from the others.
To obtain the final metallicity we decided
to use a central estimate, defined as the mean of the two central
values, which is akin to the median and more robust than the mean. In this
case the central estimate is [Fe/H]$=-0.02$ , in good agreement
with the expected value of 0.00.
Our program stars are considerably cooler and more luminous than the Sun.
We analysed the spectrum of the cool metal-poor giant
HD 190287
([Fe/H]$=-1.34$, $\rm T_{eff}=5300$),
taken with EMMI, grism \# 5 and a long-slit.
We obtain
[Fe/H] $=-1.52,-1.22,-1.23,-1.37$, for the four indices, which
yields a central estimate [Fe/H]$=-1.30$.
We compared a spectrum of one of our
program stars (SgrM 139) through the EMMI long-slit with its MOS spectrum.
The derived metallicities were the same ($\rm[Fe/H]_{MOS}=-0.05$ ;
$\rm [Fe/H]_{LS} = +0.02$) and also the radial velocities
were consistent within errors ($v_{MOS}=+147 {\rm kms^{-1}}$ ;
$v_{LS}=+124 {\rm kms^{-1}}$).
\subsection{Error estimates}
In order to estimate the random errors in our metallicity estimates we
used a Montecarlo simulation. We took a synthetic spectrum
($\rm T_{eff} = 5000 K$
[Fe/H]$=-0.5$)
to which we added Poisson noise so that S/N = 10 (i.e. the S/N of our
spectra) and estimated the metallicity from this noisy spectrum.
The mean of 10000 realizations was $\rm <[Fe/H]>= -0.56$
and the standard deviation was 0.25 dex,
we take this as our estimate of the random error.
In addition one should consider the
errors arising from uncertainties in the atmospheric parameters.
The largest error comes from the microturbulent velocity.
At this resolution we are forced to rely only on strong lines
which are very sensitive to microturbulence. The value
of $\xi$ must be assumed since there is no way to fix it from
the spectrum. The associated error is larger for the more metal-rich stars,
if the microturbulence is 1 $\rm kms^{-1}$
the metallicities
increase by $\rm \Delta([Fe/H]) = 0.24\times[Fe/H]+0.38 $.
A change in effective temperature by 150 K brings about
a change of 0.2 dex in metallicity as does a change of surface
gravity by 0.5 dex.
\subsection{Previous results and future work}
Preliminary results of this work have been already presented\cite{marb,bon}.
There are two
main differences with the results presented here.
The first is that here we use pre-tabulated synthetic indices
rather than iteratively performing spectrum synthesis for
each star and each index. The second
is that in the present work
the indices are measured with respect to the pseudo-continuum,
rather than with respect to an estimate of the true continuum, obtained
by iteratively re-normalizing the observed spectrum using the
current estimate of the best-matching synthetic spectrum, as done
previously.
The re-normalization procedure was abandoned because we
realized that it is strongly
dependent on the initial metallicity estimate. In fact it tends
to relax the observed spectrum onto the first guess. Differences
on the order of 0.2 dex are obtained from the same observed spectrum
but initial metallicity guesses differing by 0.5 dex.
Our present results are in agreement with our previous estimates
at the level of 0.1 dex except for the two most metal-poor stars
(SgrM 124 and SgrM 115)
for which we find a metallicity which is 0.32 and 0.41 dex lower.
We have so far not determined [Mg/Fe]
because the indices M1 and M2 depend on both [Mg/H] and [Fe/H].
Our grid of synthetic spectra includes only ``standard'' values
of [Mg/Fe] (0.0 down to [Fe/H]$=-0.5$ and +0.4 below). On the other
hand, our preliminary results suggest different [Mg/Fe] ratios,
especially for $\rm -0.5 < [Fe/H] < 0.0 $ .
On the contrary the F indices are essentially independent of
the abundance of $\alpha$ elements. Determining [Mg/Fe] is still
possible, once [Fe/H] is fixed through the F indices, but requires
a grid of synthetic spectra computed for a range of [Mg/Fe] values
or direct spectrum synthesis on a star by star basis. This
shall be the object of our future work
\section{Results on Sagittarius}
The metallicity distribution of our 22 candidate Sagittarus members is
shown in Figure 1. The peak at metallicity +0.5 reflects our choice
to avoid extrapolation at metallicities
higher than our most metal-rich grid point; when a star showed
indices indicative of higher metallicity we assigned to it the value +0.5.
The distribution appears to be distinctly tri-modal: there is a metal-rich
population
whose mean metallicity is yet undefined
and there are two metal-poor populations
which peak around $-1.2$ and $-0.5$ respectively.
The existence of the two metal-poor populations is in substantial
agreement with previous findings\cite{mara}.
The presence of a metal-rich
population is something new and unexpected.
It is not yet clear whether the metal-rich stars belong to Sagittarius
or are Bulge interlopers, although in the latter case their
number is uncomfortably large and may require a revision of our
understanding of Bulge kinematics.
An independent study based on Keck
High-Res spectra\cite{sme}, finds that 2 out of 7 stars, with radial
velocity compatible with Sagittarius membership,
are in fact metal-rich. Yet another photometric survey\cite{bel}
suggests the possible existence of a RGB sequence
considerably more metal-rich than 47 Tuc.
Although we are at an early stage to draw definitive conclusions,
circumstantial evidence is arising for a metal-rich
population in Sagittarius. This does not fit with
our present understanding of
chemical evolution of dwarf spheroidals, however we should keep an open mind
and stick to the observations.
\section{Perspectives for the use of VLT}
VLT will quite likely play a fundamental role
in the study of Local Group Galaxies and Sagittarius in particular.
With UVES it shall be possible
to do a detailed line by line analysis
for stars along the RGB to below the Horizontal Branch
($V_{HB}\approx 18.2$). The Turn-off $21\le V \le 21.5$ will still
be out of reach of UVES though.
New perspectives are opened by FORS. The resolution shall be
at most 1/2 of that used by us at NTT, however the indices
we have defined measure such strong features that they will be usable
even at this lower resolution, as shown
by simulations we carried out. Moreover the MOS slitlets of FORS
have sharp edges and it will be possible
to flux-calibrate the spectra, thus allowing to measure spectral
indices in absolute flux.
This will recover all the information contained the continuum.
The spectral coverage (350 - 590 nm) will allow to use
the Ca II H and K lines and probably other iron-related indices
in the range 400 - 500 nm . Finally with VLT+FORS a S/N of about 100
can be reached in one hour for stars of $V\approx 18.5$, with
this S/N our Montecarlo simulations predict
the random error in metallicity determinations to be around
0.02 dex.
The use of a 3m class telescope, such as NTT, for
preliminary observations aimed at the determination of radial velocities
has proved to be too time-consuming to be worth the effort.
FORS will be able to provide good quality spectra from which
radial velocities, metallicities and key abundance ratios, such as
[Mg/Fe] and [Ca/Fe], can be determined.
The real killer in the studies of Local Group galaxies shall
be Giraffe fed by 130 fibres in MEDUSA mode. The highest resolution
achieved by Giraffe (15000) is high enough to perform a classical
line by line analysis. However the number of spectra we shall get
(several hundreds per night) will preclude the use
of this approach. We will have to find new ways to estimate
abundances from these spectra which will be able to cope with
the data flow provided by Flames+Giraffe.
One possibility is suggested by our low-resolution experience:
define spectral indices that
can be related to abundances of particular elements and tabulate
their value from synthetic spectra.
The estimation of an abundance requires then only an interpolation
in an appropriate table.
Different indices will be appropriate for different temperature and
luminosity regimes. Now is the time to investigate
both theoretically and observationally
which features will provide useful indices.
For example some indices may be defined to measure essentially
strong iron lines others essentially weak iron lines,
the balance of the two will allow to determine microturbulence.
The index approach is not the only possibility. Other
very promising techniques such
as autocorrelation and cross-correlation should be investigated.
A whole range of methods could complement each other to extract
information from the spectra, however if any particular
star seems deserving of special attention, all the spectra will be nicely
archived and one can go back and do the good old line by line analysis.
\section{Acknowledgements}
I wish to thank G. Marconi, P. Molaro and L. Pasquini for allowing
me to present results of common work in advance of publication.
This work is
based on observations collected at the European Southern Observatory, Chile,
ESO N$^\circ$ 57.E-0586.
\addcontentsline{toc}{section}{References}
|
\section{Introduction}
Quantum field theory is undoubtedly a very successful theory except
for its mathematically rigorous foundation. The basic quantities of
quantum field theory are quantum fields, generically denoted by $\varphi(x)$,
which are operator-valued singular functions of $N$-dimensional
space-time $x^\mu$. In the standard formulation, their properties are
governed by the action $S$, which is an $N$-dimensional integral of
the Lagrangian density ${\cal L}$ (a function of quantum fields at the
same space-time point). Field equations and canonical (anti)commutation
relations are derived from $S$ by the standard procedure. Equal-time
(anti)commutation relations follow from the canonical ones.
\par
Field equations and equal-time (anti)commutation relations define the
operator properties of quantum fields, that is, it is supposed to
determine the operator algebra of quantum fields. Next, this operator
algebra is represented in terms of state vectors, whose totality forms
an infinite dimensional complex linear space equipped with (indefinite)
inner product. This way of formulating the theory is operator-formalism
approach. Evidently, this way of thinking is most natural as the
formulation of quantum fields. Nevertheless, this approach has never
been seriously considered as the standard approach to quantum field
theory. The main reasons for this are as follows. First, it has been
unknown how to find the solution in this approach. Second, it has been
unknown how to deal with the divergence problem which arises from the
singular nature of quantum fields.
\par
The present-day's standard approach is covariant perturbation theory.
It is based on the action $S$, rather than field equations. Decomposing
it artificially into its free part and its interaction part, one obtains
all Green's functions explicitly in terms of Feynman diagrams.
Furthermore, renormalization theory tells us how to deal with
the divergence problem: divergent contributions are absorbed into
counter terms.
\par
Path-integral formalism gives us the generating functional, $Z$, of the
Green's functions. If path integral is defined as the infinite multiple
integrals with respect to the expansion coefficients of quantum fields
in terms of their free wave functions, $Z$ reproduces the perturbative
expressions. If path integral is defined by using an expansion in terms
of the functions other than free wave functions, it is not clear whether
or not $Z$ yields the Green's functions which reproduce the perturbative
ones (if expanded), but it is usually supposed that there exists
an appropriate path-integral measure giving the expected results.
Although we cannot accept this assumption without
reservation,\footnote{This problem is particularly relevant in quantum
gravity because gravity's free wave functions are rather artificial
quantities. For example, the sum over all possible manifolds having
a particular metric signature cannot be realized by a simple infinite
multiple integral with respect to expansion coefficients because
a definite metric signature requires to insert a product of $\theta$
functions of the determinant and principal minors of $g_{\mu\nu}$ into
the path integral.} we do not strictly distinguish perturbative
approach and path-integral one in the present paper.
\par
By taking the logarithm of $Z$, one obtains the generating functional,
$W$, of connected Green's functions. Moreover, the effective action
$\Gamma$, which is the generating functional of amputated proper
Feynman diagrams, is obtained as the functional Legendre transform
of $W$. The renormalization is neatly carried out in $\Gamma$ and
therefore the anomaly problem is usually discussed also in $\Gamma$.
It should be noted here that perturbation series is not unique because
the decomposition of the action into its free part and its interaction
part is generally altered by a nonlinear redefinition of fields.
Accordingly, the effective action $\Gamma$ is a quantity which generally
changes under the redefinition of fields. Thus, in the path-integral
approach, neither renormalization procedure nor the anomaly problem are
quite independent of the choice of quantum fields. This point becomes
crucial when unphysical fields, whose natural definitions are not
necessarily unique, play important roles as in quantum gravity.
\par
Now, we have recently succeeded in formulating the method of finding
the solution in the operator-formalism approach.\cite{AN1}
Our method is as follows. From the field equations and equal-time
(anti)commutation relations, we explicitly construct all independent
$N$-dimensional (anti)commutation relations, by expanding them, if
necessary, into the power series with respect to the parameters
involved. We then calculate independent $N$-dimensional multiple
(anti)commutators. The representation of the field algebra in terms
of state vectors is constructed by giving all $n$-point Wightman
functions ($n=1,\,2,\,\ldots\,$), i.e., vacuum expectation values of
simple products of $n$ quantum fields, $\varphi_1(x_1)\varphi_2(x_2)\cdots
\varphi_n(x_n)$, so as to be consistent with the $(n-1)$ple
(anti)commutators and with the energy positivity
conditions.\footnote{The Wightman function is a boundary value of
an analytic function of the variables $x_i{}^0-x_j{}^0$ $\ (i<j)$
from the lower half-planes.}
Here, in contrast with the axiomatic field theory,\cite{SW} we need the
Wightman functions involving composite fields, where a composite field
is a product of fields at the same space-time point. When we set some
of space-time points coincident in a higher-point Wightman function,
we generally encounter divergent terms, which must be simply discarded
in such a way that the resultant be independent of the ordering of
the constituent fields of the composite field (``generalized normal
product rule''). In this procedure, we do not introduce anything like
a counter term. This is because a well-defined representation of the
field algebra should be free of divergence; our standpoint is similar
to that of the Lehmann-Symanzik-Zimmermann formalism,\cite{LSZ} in which
they developed the renormalized perturbation theory without encountering
any unrenormalized quantities.
\par
Of course, it is extremely difficult to carry out our way of finding
the solution in realistic models. But, fortunately, we can explicitly
construct the exact solutions in some two-dimensional models by our
method.\cite{AN2,AN3,AN4,AN5,Ikeda}
Our results are seen to be quite satisfactory, but we encounter an anomalous
phenomenon, which we call ``field-equation anomaly'', in quantum-gravity
models\cite{AN2,AN3,AN4,AN5} (but not in gauge-theory models\cite{Ikeda}):
By construction, our Wightman functions are consistent with all
two-dimensional (anti)commutators but not necessarily consistent with
nonlinear field equations because there we encounter products of fields
at the same space-time point.
In any of the quantum-gravity models which we have exactly solved so far,
one of field equations is {\it slightly\/} violated at the level of
representation.\ifundefined{abstsize}\setcounter{footn}{0}\fi\footnote{The violation is
{\it slight\/} in the sense that an anomaly-free equation can be obtained
by differentiating the original field equation once or twice.}
This is the field-equation anomaly. It is different from the
conventional anomalies which arise in connection with particular
symmetries. Rather, as clarified in our previous work,\cite{AN6}
various conventional anomalies\cite{AN7} are systematically explained
on the basis of the field-equation anomaly and their ambiguities are
shown to be caused by the nonuniqueness of perturbation theory.
In this sense, we regard the field-equation anomaly as a more fundamental
concept.
\par
The purpose of the present paper is to make comparison between the
solution obtained by the perturbative approach and the one obtained
by the operator-formalism approach explicitly in the conformal-gauge
two-dimensional quantum gravity. The exact solution of this model
obtained previously\cite{AN4} can be written in terms of {\it tree
diagrams only}. If the same were true also in the perturbative
solution, no anomaly could be present. As is well known, however,
this model has the conformal anomaly except for $D=26$, where $D$
denotes the number of the scalar fields which can be interpreted as
the string coordinates. We trace the cause of this paradox and find
that the perturbative approach induces some one-loop Feynman diagrams,
which would not exist unless a nonzero contribution arose from a zero
field. The cause of this strange phenomenon is found to be the use of
T$^*$-product (covariantized T-product) of quantum
fields\ifundefined{abstsize}\setcounter{footn}{1}\fi\footnote{T$^*$-product is
a T-product modified in such a way tha
\ifundefined{abstsize}
\hbox{$\,\partial_1{}\!^0\hbox{T}^{\hskip-.5pt*}\hskip-.5pt\varphi_1(x_1)
\cdots\varphi_n(x_n)
\!=\!\hbox{T}^{\hskip-.5pt*}\hskip-.5pt\partial_1{}\!^0\varphi_1(x_1)
\cdots\varphi_n(x_n)$.}
\else
{} $\partial_1{}^0\hbox{T}^*\varphi_1(x_1)\cdots\varphi_n(x_n)
=\hbox{T}^*\partial_1{}^0\varphi_1(x_1)\cdots\varphi_n(x_n)$.
\fi
}
in the perturbative or path-integral approach. More generally, in
the present paper, we clarify that various anomalous behaviors of
this model found in the perturbative approach are caused by the use
of T$^*$-product.
\par
In the present paper, we compare the perturbative or path-integral
approach with the operator-formalism approach in the conformal-gauge
two-dimensional quantum gravity,\ifundefined{abstsize}\break\else{} \fi whose
Lagrangian density is given in \S2. In \S3, \S4 and \S5, respectively,
we discuss this model by the operator-formalism approach, by the
perturbative approach and by the path-integral approach.
In \S6, we criticize the so-called ``FP-ghost number current anomaly''.
The final section is devoted to discussions.
\ifundefined{abstsize}
\newpage
\fi
\section{Preliminaries}
Throughout the present paper, we consider the BRS formalism of the
conformal-gauge two-dimensional quantum gravity, in which the conformal
degree of freedom is already eliminated.\cite{Yang} Quantum fields are
contravariant tensor-density gravitational field $\tg^{\mu\nu}$, FP
ghost $c^\mu$, FP anti-ghost ${\skew3\bar c}\hspace*{1pt}_{\mu\nu}$, B field $\,\tilde{\! b}{}_{\mu\nu}$
and scalar fields $\phi_M\; (M=0,\,1,\,\ldots,\,D-1)$.
Their BRS transforms are as follows:
\begin{eqnarray}
{\mbf\delta_*}\, \tg^{\mu\nu}&=& \tg^{\mu\sigma}\partial_\sigma c^\nu
+\tg^{\nu\sigma}\partial_\sigma c^\mu
-\partial_\sigma(\tg^{\mu\nu}c^\sigma), \\
{\mbf\delta_*}\, c^\mu &=&-c^\sigma\partial_\sigma c^\mu, \\
{\mbf\delta_*}\, {\skew3\bar c}\hspace*{1pt}_{\mu\nu} &=& i\,\tilde{\! b}{}_{\mu\nu}, \\
{\mbf\delta_*}\, \,\tilde{\! b}{}_{\mu\nu} &=&0,\\
{\mbf\delta_*}\, \phi_M &=& -c^\sigma\partial_\sigma\phi_M.
\end{eqnarray}
Since $\tg^{\mu\nu}$ has only two degrees of freedom because det$\tg^{\mu\nu}
=-1$ , it is parametrized as
\begin{eqnarray}
\tg^{\mu\nu}&=&(\eta^{\mu\nu}+h^{\mu\nu})(1-\det{h^{\sigma\tau}})^{-1/2},
\label{eq:2,6}
\end{eqnarray}
where $h^{\mu\nu}$ is symmetric and traceless ($\eta_{\mu\nu}h^{\mu\nu}=0$).
Correspondingly, ${\skew3\bar c}\hspace*{1pt}_{\mu\nu}$ and $\,\tilde{\! b}{}_{\mu\nu}$ are also symmetric
and traceless. It is convenient to rewrite any traceless symmetric tensor
$X_{\mu\nu}$ into a vector-like quantity $X^\lambda$ by
\begin{eqnarray}
X^\lambda &=& {1\over\sqrt{2}}\xi^{\lambda\mu\nu}X_{\mu\nu}, \label{eq:2,7}
\end{eqnarray}
where $\xi^{\lambda\mu\nu}=1$ for $\lambda+\mu+\nu=$even, $=0$ otherwise.
According to \eqno(2,7), we introduce $h_\lambda$, ${\skew3\bar c}\hspace*{1pt}^\lambda$ and
$\,\tilde{\! b}{}^\lambda$.
\par
The BRS-invariant action $S=\int d^2x\,{\cal L}$ is given by the Lagrangian
density
\begin{eqnarray}
{\cal L}&=&{1\over\,2\,}\tg^{\mu\nu}\partial_\mu\phi_M\cdot\partial_\nu\phi^M
-{1\over\,2\,}\,\tilde{\! b}{}^\lambda h_\lambda
-{i\over\,2\,}{\skew3\bar c}\hspace*{1pt}^\lambda{\mbf\delta_*}\,(h_\lambda), \label{eq:2,8}
\end{eqnarray}
where $\tg^{\mu\nu}$ is given by \eqno(2,6), that is,
\begin{eqnarray}
{\cal L}&=&{\cal L}_0+\lag_{\hbox{\tiny I}}, \label{eq:2,9}
\end{eqnarray}
with
\begin{eqnarray}
{\cal L}_0&=& {1\over\,2\,}\eta^{\mu\nu}\partial_\mu\phi_M\cdot\partial_\nu\phi^M
-{1\over\,2\,}\,\tilde{\! b}{}^\lambda h_\lambda
-{i\over\sqrt{2}}\xi_{\lambda\mu\nu}{\skew3\bar c}\hspace*{1pt}^\lambda\partial^\mu c^\nu,
\label{eq:2,10} \\
\lag_{\hbox{\tiny I}}&=& h_\lambda
\bigg({1\over2\sqrt{2}}\xi^{\lambda\mu\nu}\partial_\mu\phi_M\cdot
\partial_\nu\phi^M
\if@preprint
-{i\over\,2\,}\xi^{\lambda\mu\nu}\xi_{\rho\mu\sigma}{\skew3\bar c}\hspace*{1pt}^\rho
\partial_\nu c^\sigma
-{i\over\,2\,}\partial_\sigma{\skew3\bar c}\hspace*{1pt}^\lambda\cdot c^\sigma\bigg)
+O(h^2). \label{eq:2,11} \hspace*{20pt}
\else
\!-\!{i\over\,2\,}\xi^{\lambda\mu\nu}\xi_{\rho\mu\sigma}{\skew3\bar c}\hspace*{1pt}^\rho
\partial_\nu c^\sigma
\!-\!{i\over\,2\,}\partial_\sigma{\skew3\bar c}\hspace*{1pt}^\lambda\cdot c^\sigma\bigg)
\!+\!O(h^2). \label{eq:2,11} \hspace*{40pt}
\fi
\end{eqnarray}
The higher-order terms $O(h^2)$ are unnecessary to be specified because they
contribute neither to field equations nor to canonical (anti)commutation
relations. Furthermore, they give no contribution to perturbation theory.
Thus, we may discard them.
\par
It should be noted that the action is invariant under the FP-ghost
conjugation\footnote{So far, this fact has been overlooked because the
FP antighost was treated as a tensor.}
\begin{eqnarray}
&&c^\lambda \ \longrightarrow \ {\skew3\bar c}\hspace*{1pt}^\lambda, \nonumber \\
&&{\skew3\bar c}\hspace*{1pt}^\lambda \ \longrightarrow \ c^\lambda, \nonumber \\
&&\,\tilde{\! b}{}^\lambda \ \longrightarrow \
\,\tilde{\! b}{}^\lambda - i\xi^{\lambda\mu\nu}\xi_{\rho\mu\sigma}{\skew3\bar c}\hspace*{1pt}^\rho
\partial_\nu c^\sigma
- i\partial_\sigma{\skew3\bar c}\hspace*{1pt}^\lambda\cdot c^\sigma
\nonumber \\
&&\hspace*{55pt}
+ i\xi^{\lambda\mu\nu}\xi_{\rho\mu\sigma}c^\rho
\partial_\nu{\skew3\bar c}\hspace*{1pt}^\sigma
+ i\partial_\sigma c^\lambda\cdot {\skew3\bar c}\hspace*{1pt}^\sigma
\ + \ O(h),
\end{eqnarray}
as in the de Donder-gauge case.
\par
Analysis can be much simplified by introducing light-cone coordinates
$x^\pm=$
$(x^0\pm x^1)/\sqrt{2}$, because then $\xi_{\mu\nu\lambda}=0$ except
$\xi_{+++}=\xi_{---}=\sqrt{2}$. From \eqno(2,10) and \eqno(2,11) ($O(h^2)$
is omitted), we have
\begin{eqnarray}
&&{\cal L}_0=\partial_+\phi_M\cdot\partial_-\phi^M
+\Big[-{1\over\,2\,}\,\tilde{\! b}{}^+ h_+ -i{\skew3\bar c}\hspace*{1pt}^+\partial_-c^+
+ \quad (\ +\ \longleftrightarrow\ -\ )\Big],
\label{eq:2,13} \\
&&\lag_{\hbox{\tiny I}}=h_+\Big[{1\over\,2\,}\partial_+\phi_M\cdot\partial_+\phi^M
- i{\skew3\bar c}\hspace*{1pt}^+\partial_+c^+
-{i\over\,2\,}\Big(\partial_+{\skew3\bar c}\hspace*{1pt}^+\cdot c^+
+\partial_-{\skew3\bar c}\hspace*{1pt}^+\cdot c^- \Big) \Big] \
\nonumber \\
&&\hspace*{50pt}
+ \quad (\ +\ \longleftrightarrow\ -\ ).
\label{eq:2,14}
\end{eqnarray}
In subsequent sections, we start with \eqno(2,9) together with \eqno(2,13)
and \eqno(2,14). For later convenience, we introduce the following notation.
\begin{eqnarray}
\skew3\widetilde{\calT}^\pm &\equiv& \partial_\pm\phi_M\cdot\partial_\pm\phi^M
-2i{\skew3\bar c}\hspace*{1pt}^\pm\partial_\pm c^\pm
-i\partial_\pm{\skew3\bar c}\hspace*{1pt}^\pm\cdot c^\pm, \\
\skew3\widetilde{\tcalT}^\pm &\equiv& \skew3\widetilde{\calT}^\pm - i\partial_\mp{\skew3\bar c}\hspace*{1pt}^\pm\cdot c^\mp.
\end{eqnarray}
Then we have
\begin{eqnarray}
{\partial \lag_{\hbox{\tiny I}} \over \partial h_\pm} &=& {1\over\,2\,}\skew3\widetilde{\tcalT}^\pm.
\end{eqnarray}
\par
The Noether currents of the BRS invariance and the FP-ghost number
conservation are given by
\begin{eqnarray}
j_b{}^\mp &=& -c^\pm \partial_\pm \phi_M\cdot \partial_\pm \phi^M
-i{\skew3\bar c}\hspace*{1pt}^\pm c^\pm \partial_\pm c^\pm,
\label{eq:2,18} \\
j_c{}^\mp &=& -i{\skew3\bar c}\hspace*{1pt}^\pm c^\pm,
\end{eqnarray}
respectively. They are of course conserved.
\ifundefined{abstsize}
\newpage
\fi
\section{Operator-formalism Approach}
For the sake of comparison, we briefly review our previous results of the
exact solution obtained by the operator-formalism approach.\cite{AN4}
\par
The field equations are as follows:
\begin{eqnarray}
&&h_\pm=0, \label{eq:3,1} \\
&&\,\tilde{\! b}{}^\pm=\skew3\widetilde{\calT}^\pm, \label{eq:3,2} \\
&&\partial_\mp X^\pm=0 \quad \hbox{for } X^\pm=c^\pm,\ {\skew3\bar c}\hspace*{1pt}^\pm,\ \,\tilde{\! b}{}^\pm,
\label{eq:3,3} \\
&&\partial_+\partial_-\phi_M=0, \label{eq:3,4}
\end{eqnarray}
where \eqno(3,3) for $X^\pm=\,\tilde{\! b}{}^\pm$ is derived by differentiating \eqno(3,2).
Note that $c^\pm,\ {\skew3\bar c}\hspace*{1pt}^\pm,\ \,\tilde{\! b}{}^\pm$ and $\partial_\pm\phi_M$ are the
functions of a single variable $x^\pm$ only.
\par
Canonical quantization is carried out by taking $\phi_M$ and $c^\pm$ only as
the canonical variables. Since they are free fields, their nonvanishing %
two-dimensional (anti)\-commutators
are easily obtained; we have
\begin{eqnarray}
&&[\partial_\pm\phi_M(x),\;\phi^N(y)]
=-{i\over\,2\,}\delta_M{}^N \delta(x^\pm-y^\pm), \label{eq:3,5} \\
&&\{c^\pm(x),\;{\skew3\bar c}\hspace*{1pt}^\pm(y)\}=-\delta(x^\pm-y^\pm).
\end{eqnarray}
The commutation relations involving $\,\tilde{\! b}{}^\pm$ are calculated by using
\eqno(3,2):
\begin{eqnarray}
{[} \,\tilde{\! b}{}^\pm(x),\;\phi_M(y) ]
&=&-i\partial_\pm\phi_M(x)\cdot\delta(x^\pm-y^\pm), \\
{[} \,\tilde{\! b}{}^\pm(x),\;c^\pm(y) ]
&=&-ic^\pm(x)\delta'(x^\pm-y^\pm)
-2i\partial_\pm c^\pm(x)\cdot\delta(x^\pm-y^\pm), \\
{[} \,\tilde{\! b}{}^\pm(x),\;{\skew3\bar c}\hspace*{1pt}^\pm(y) ]
&=& 2i{\skew3\bar c}\hspace*{1pt}^\pm(x)\delta'(x^\pm-y^\pm)
+i\partial_\pm{\skew3\bar c}\hspace*{1pt}^\pm(x)\cdot\delta(x^\pm-y^\pm), \label{eq:3,9} \\
{[} \,\tilde{\! b}{}^\pm(x),\;\,\tilde{\! b}{}^\pm(y) ]
&=&i[\,\tilde{\! b}{}^\pm(x)+\,\tilde{\! b}{}^\pm(y)]\delta'(x^\pm-y^\pm). \label{eq:3,10}
\end{eqnarray}
Evidently, \eqno(3,10) is the BRS transform of \eqno(3,9).
\par
Since no new operators are encountered in the right-hand sides of
\eqno(3,5)-\eqno(3,10), we can easily calculate all multiple
(anti)commutators explicitly.
We then construct all truncated\footnote{Truncation means to drop the
contributions from vacuum intermediate states. The {\it truncated\/} Wightman
function corresponds to the {\it connected\/} Green's function.
In the present model, the distinction between truncated and nontruncated
appears only for $n\geqq4$.} Wightman functions so as to be consistent with
all multiple (anti)commutators under the energy positivity condition.
\par
The 1-point functions are, in principle, completely arbitrary. But we set
all of them equal to zero because we should not deliberately violate any of
\eqno(3,1), FP-ghost number conservation, BRS invariance and $O(D)$ symmetry.
\par
The nonvanishing {\it truncated\/} $n$-point Wightman functions are those
which consist of $(n-2)$ $\,\tilde{\! b}{}^\pm$'s and of either $c^\pm$ and ${\skew3\bar c}\hspace*{1pt}^\pm$
or two $\phi_M$'s. Diagrammatically, they are represented by tree diagrams.
Although we have explicitly constructed all of them,\cite{AN4} we here
quote 2-point and 3-point ones only.
\par
Nonvanishing 2-point Wightman functions are
\begin{eqnarray}
&&\wightman{\phi_M(x_1)\phi^N(x_2)}
=\delta_M{}^N D^{\hbox{\tiny (+)}}(x_1-x_2),\\
&&\wightman{c^\pm(x_1){\skew3\bar c}\hspace*{1pt}^\pm(x_2)}
=\wightman{{\skew3\bar c}\hspace*{1pt}^\pm(x_1)c^\pm(x_2)}
=-2i\partial_\pmD^{\hbox{\tiny (+)}}(x_1-x_2),
\end{eqnarray}
where\footnote{$D^{\hbox{\tiny (+)}}(x)$ itself is infrared divergent and therefore
requires the introduction of infrared \linebreak[3]
cutoff.~\cite{Nakanishi}}
\begin{eqnarray}
&& \partial_\pmD^{\hbox{\tiny (+)}}(x)\equiv -{1\over4\pi}\cdot{1\over x^\pm-i0}.
\end{eqnarray}
Nonvanishing 3-point ones are
\begin{eqnarray}
&&\wightman{\phi_M(x_1)\,\tilde{\! b}{}^\pm(x_2)\phi^N(x_3)}
=-2\delta_M{}^N \partial_\pmD^{\hbox{\tiny (+)}}(x_1-x_2)\cdot\partial_\pmD^{\hbox{\tiny (+)}}(x_2-x_3),
\hspace*{30pt}
\label{eq:3,14} \\
&&\wightman{c^\pm(x_1)\,\tilde{\! b}{}^\pm(x_2){\skew3\bar c}\hspace*{1pt}^\pm(x_3)}
=8i\partial_\pmD^{\hbox{\tiny (+)}}(x_1-x_2)\cdot\partial_\pm{}^2D^{\hbox{\tiny (+)}}(x_2-x_3) \nonumber\\
&&\hspace*{120pt}
-4i\partial_\pm{}^2D^{\hbox{\tiny (+)}}(x_1-x_2)\cdot\partial_\pmD^{\hbox{\tiny (+)}}(x_2-x_3)
\label{eq:3,15}
\end{eqnarray}
and their permutated ones, whose expressions are obtained from the above
by changing some of $D^{\hbox{\tiny (+)}}(x)$'s into $-[D^{\hbox{\tiny (+)}}(x)]^*$ so as to become consistent
with the energy-positivity condition (and by changing the overall sign if
the order of $c^\pm$ and ${\skew3\bar c}\hspace*{1pt}^\pm$ is reversed).
\par
Our system of Wightman functions is, of course, consistent with the field
algebra defined by \eqno(3,5)-\eqno(3,10). It is also consistent with
the BRS invariance and the FP-ghost number conservation. It should be noted
that we need the use of the generalized normal-product rule to check the
BRS invariance. Our system of Wightman functions is also consistent with
all {\it linear\/} field equations \eqno(3,1), \eqno(3,3) and \eqno(3,4),
but {\it not\/} with the {\it nonlinear\/} field equation \eqno(3,2).
Indeed, by using the generalized normal-product rule, we can show that
\begin{eqnarray}
\wightman{\,\tilde{\! b}{}^\pm(x_1)\,\tilde{\! b}{}^\pm(x_2)}&=&0, \\
\wightman{\,\tilde{\! b}{}^\pm(x_1)\skew3\widetilde{\calT}^\pm(x_2)}
&=&\wightman{\skew3\widetilde{\calT}^\pm(x_1)\skew3\widetilde{\calT}^\pm(x_2)} \nonumber\\
&=&2(D-26)[\partial_\pm{}^2D^{\hbox{\tiny (+)}}(x_1-x_2)]^2
\end{eqnarray}
in contradiction with \eqno(3,2). Thus the field equation \eqno(3,2),
{\it modulo\/} \eqno(3,3) for $X^\pm=\,\tilde{\! b}{}^\pm$, is violated at the level
of the representation in terms of state vectors.
We call this matter ``field-equation anomaly''. This phenomenon is
encountered also in several two-dimensional quantum-gravity
models.\cite{AN2,AN3,AN4,AN5}
\par
The BRS Noether current \eqno(2,18) can be rewritten as
\begin{eqnarray}
&&j_b{}^\mp=j_b'{}^\mp + (\,\tilde{\! b}{}^\pm-\skew3\widetilde{\calT}^\pm)c^\pm, \\
\noalign{\noindent with}
&&j_b'{}^\mp \equiv -\,\tilde{\! b}{}^\pm c^\pm + i{\skew3\bar c}\hspace*{1pt}^\pm c^\pm \partial_\pm c^\pm.
\end{eqnarray}
At the operator level, $j_b'{}^\mp$ strictly equals $j_b{}^\mp$.
But this equality no longer holds at the representation level because of
the appearance of the field-equation anomaly. Indeed, $j_b{}^\mp$ is
anomalous for $D\not=26$, while $j_b'{}^\mp$ is free of anomaly for
any value of $D$.\cite{AN4,AN5} On the other hand, $j_c{}^\mp$ is free
of anomaly without making any modification.
\section{Perturbative approach}
The perturbative approach is so familiar to everybody that no explanation
about it is necessary. Nevertheless, when compared with the
operator-formalism approach, the perturbative approach is seen to yield
some surprising results.
\par
The Lagrangian density ${\cal L}$ is decomposed into the free one ${\cal L}_0$,
which is quadratic with respect to the fields adopted as the basic ones,
and the remainder $\lag_{\hbox{\tiny I}}$, called the interaction Lagrangian density.
\par
The Feynman propagators are obtained by taking the inverse of the
differential operator sandwiched by the fields in ${\cal L}_0$.
Thus, from \eqno(2,13), we have the following nonvanishing Feynman
propagators:\footnote{The subscript 0 indicates that the propagators are
free ones.}
\begin{eqnarray}
&&\taustar{\phi_M(x_1)\phi^N(x_2)}_0=\delta_M{}^ND_{\hbox{\tiny F}}(x_1-x_2),
\label{eq:4,1}\\
&&\taustar{\,\tilde{\! b}{}^\pm(x_1)h_\pm(x_2)}_0=-2i\delta^2(x_1-x_2),
\label{eq:4,2}\\
&&\taustar{{\skew3\bar c}\hspace*{1pt}^\pm(x_1)c^\pm(x_2)}_0=-2i\partial_\pmD_{\hbox{\tiny F}}(x_1-x_2),
\label{eq:4,3}
\end{eqnarray}
where
\begin{eqnarray}
\partial_\pmD_{\hbox{\tiny F}}(x)&\equiv&\theta(x^0)\partial_\pmD^{\hbox{\tiny (+)}}(x)
+\theta(-x^0)\partial_\pm[D^{\hbox{\tiny (+)}}(x)]^*
\nonumber\\
&=&-{1\over 4\pi}\bigg[ {\theta(x^\mp+x^\pm)\over x^\pm-i0}
+{\theta(-x^\mp-x^\pm)\over x^\pm+i0} \bigg].
\label{eq:4,4}
\end{eqnarray}
\par
It is quite remarkable that $\taustar{\,\tilde{\! b}{}^\pm\,h_\pm}_0$ is
{\it nonvanishing\/} in spite of the fact that $h_\pm$ is a {\it zero\/}
operator as is seen from \eqno(3,1). In contrast with the Wightman
functions, the T$^*$-product does not respect the validity of the
field equations. As is seen from \eqno(4,3), it is also inadmissible to
set $\partial_\mp c^\pm=\partial_\mp{\skew3\bar c}\hspace*{1pt}^\pm=0$ in the perturbative
approach. Hence we cannot discard the terms involving
$\partial_\mp{\skew3\bar c}\hspace*{1pt}^\pm$ in $\lag_{\hbox{\tiny I}}$, that is, we have to distinguish
$\skew3\widetilde{\tcalT}^\pm$ from $\skew3\widetilde{\calT}^\pm$. Thus the beautiful result of the operator
formalism that $c^\pm$, ${\skew3\bar c}\hspace*{1pt}^\pm$, $\,\tilde{\! b}{}^\pm$ and $\partial_\pm\phi_M$
are irrelevant to $x^\mp$ is no longer valid in the perturbative approach.
This fact makes the perturbative calculation complicated and sometimes
misleading, as we shall see later.
\par
By using $\lag_{\hbox{\tiny I}}$ given by \eqno(2,14), we can easily calculate the
$n$-point Green's functions. For example, we have
\begin{eqnarray}
&&\taustar{\phi_M(x_1)\,\tilde{\! b}{}^\pm(x_2)\phi^N(x_3)}
=-2\delta_M{}^N\partial_\pmD_{\hbox{\tiny F}}(x_1-x_2)\cdot\partial_\pmD_{\hbox{\tiny F}}(x_2-x_3),
\label{eq:4,5} \\
&&\taustar{c^\pm(x_1)\,\tilde{\! b}{}^\pm(x_2){\skew3\bar c}\hspace*{1pt}^\pm(x_3)}
=8i\partial_\pmD_{\hbox{\tiny F}}(x_1-x_2)\cdot\partial_\pm{}^2D_{\hbox{\tiny F}}(x_2-x_3)
\nonumber \\
&&\hspace*{135pt}
-4i\partial_\pm{}^2D_{\hbox{\tiny F}}(x_1-x_2)\cdot\partial_\pmD_{\hbox{\tiny F}}(x_2-x_3),
\label{eq:4,6}\\
&&\taustar{c^\pm(x_1)\,\tilde{\! b}{}^\pm(x_2){\skew3\bar c}\hspace*{1pt}^\mp(x_3)}
=-2\delta^2(x_1-x_2)\partial_\mpD_{\hbox{\tiny F}}(x_2-x_3). \label{eq:4,7}
\end{eqnarray}
Evidently, \eqno(4,5) and \eqno(4,6) correspond to \eqno(3,14) and to
\eqno(3,15), respectively. However, \eqno(4,7) is a result peculiar to
the T$^*$-product. This result is seen to be consistent with the
Ward-Takahashi identity
\begin{eqnarray}
&&\taustar{{\mbf\delta_*}\,(c^\pm(x_1){\skew3\bar c}\hspace*{1pt}^\pm(x_2){\skew3\bar c}\hspace*{1pt}^\mp(x_3))}=0,
\end{eqnarray}
because the second term of
${\mbf\delta_*}\, c^\pm=-c^\pm\partial_\pm c^\pm-c^\mp\partial_\mp c^\pm$
contributes.
\par
Now, we come to the crucial point. In sharp contrast with the case of
the operator-formalism approach, the perturbative approach yields quite
a nontrivial result for the $n$-point Green's function consisting of
B-fields only. Indeed, its connected part is given by a sum over
one-loop Feynman diagrams.
For example, we consider $\taustar{\,\tilde{\! b}{}^\lambda(x_1)\,\tilde{\! b}{}^\rho(x_2)}$.
Because of the nonvanishing of \eqno(4,2), the second-order perturbation
term yields
\begin{eqnarray}
&&\taustar{\,\tilde{\! b}{}^\lambda(x_1)\,\tilde{\! b}{}^\rho(x_2)}
=\taustar{\skew3\widetilde{\tcalT}^\lambda(x_1)\skew3\widetilde{\tcalT}^\rho(x_2)}_0. \label{eq:4,9}
\end{eqnarray}
Therefore, we have
\begin{eqnarray}
&&\taustar{\,\tilde{\! b}{}^\pm(x_1)\,\tilde{\! b}{}^\pm(x_2)}
=2(D-26)[\partial_\pm{}^2D_{\hbox{\tiny F}}(x_1-x_2)]^2,
\label{eq:4,10} \\
&&\taustar{\,\tilde{\! b}{}^\pm(x_1)\,\tilde{\! b}{}^\mp(x_2)}=-{D-2\over2}[\delta^2(x_1-x_2)]^2.
\label{eq:4,11}
\end{eqnarray}
They are divergent and therefore require the introduction of counter terms.
Note that the use of the T$^*$-product is responsible for the appearance
of these divergences.
\par
The nonvanishing of the Green's functions consisting of B-fields only
implies the violation of the BRS invariance.
In the de Donder gauge case, Takahashi\cite{Takahashi} proposed to
convert the violation of the BRS invariance for $D\not=26$ in the
two-point B-field Green's function into the conformal
anomaly.\footnote{He made no mention
about how to remove the BRS violation in the {\it higher-point\/} functions.}
We apply his line of thought to the present model.
In addition to \eqno(4,10) and \eqno(4,11), we must take it into account
the following exact two-point Green's functions:
\begin{eqnarray}
&&\taustar{h_\lambda(x_1)h_\rho(x_2)}=0, \\
&&\taustar{\,\tilde{\! b}{}^\lambda(x_1)h_\rho(x_2)}
=-2i\delta_\rho{}^\lambda\delta^2(x_1-x_2). \label{eq:4,13}
\end{eqnarray}
The two-point functions of the effective action $\Gamma$ is obtained by
taking the matrix inverse of \eqno(4,10)-\eqno(4,13).
Accordingly, we have
\begin{eqnarray}
\Gamma&=&\int d^2x_1\int d^2x_2\bigg[
-{1\over\,2\,}\delta^2(x_1-x_2)\,\tilde{\! b}{}^\lambda(x_1)h_\lambda(x_2) \nonumber\\
&& \qquad
+{D-26\over2}i\sum_{\alpha=\pm}
([\partial_\alpha{}^2D_{\hbox{\tiny F}}(x_1-x_2)]^2)_{\hbox{\tiny R}}
h_\alpha(x_1)h_\alpha(x_2)
+ \ \cdots\ \bigg], \label{eq:4,14}
\ifundefined{abstsize} \else \hspace*{30pt}\fi
\end{eqnarray}
where a subscript R indicates regularization. The BRS-violating term
in \eqno(4,14) is converted into the conformal-anomaly term by adding
the conformal degree of freedom. We do not work out this procedure
in detail because it is not our aim to do so.
\par
The important point is the violation of the BRS invariance in the B-field
Green's functions. In the de Donder gauge case, the BRS violation has
arisen by applying the dimensional regularization only to internal lines
but {\it not to external lines}.\cite{AN6,AN8}
In the present model, external lines are absent because \eqno(4,2) is
local. Instead, as is seen from \eqno(4,9), perturbative approach makes
use of the field-equation anomaly without being aware of this fact.
\section{Path-integral approach}
The path integral directly gives us the generating function of the Green's
functions, that is, it deals with the T$^*$-product quantities only.
The path integral $Z$ is formally expressed as
\begin{eqnarray}
&&Z(J)=\int \big(\prod_i{\cal D}\varphi_i\big)\,
\exp\, i\!\int d^N\!{}x({\cal L}+\sum_iJ_i\varphi_i)
\end{eqnarray}
with $Z(0)=1$, where $J_i$ denotes the source function corresponding to
the field $\varphi_i$.
\par
It is possible to derive the path-integral formula [corrected by
the Lee-Yang term proportional to $\delta^N(0)$] from the canonical
operator formalism.\cite{Buchbinder}
In this sense, the path-integral formalism can be regarded as the one
equivalent to the operator formalism. But, one should note that,
in this derivation, one must use the {\it field equations at the
representation level}.
This fact implies that the path-integral formalism cannot take care
of the existence of the field-equation anomaly.
\par
From the successful experience of discussing the anomaly problem in
gauge theories, it has been customary to believe that any anomaly always
arises from the non-invariance of the path-integral measure under
the symmetry which leaves the action $S$ invariant.
But we point out that anomalies can arise also from the field-equation
anomaly which is beyond the scope of the path-integral formalism.
\par
Let $F(\varphi)$ be an arbitrary function of $\varphi_i$'s.
The path-integral measure is supposed to be invariant under the functional
translation $\varphi_i \longrightarrow \varphi_i +\delta\varphi_i$.
Hence, by considering a variation of a field $\varphi_i$ in
$F(i^{-1}\partial/\partial J)Z|_{J=0}$, we obtain
\begin{eqnarray}
&&i\taustar{F(\varphi){\delta S\over \delta \varphi_i}}
+\taustar{{\delta F(\varphi)\over \delta \varphi_i}} =0. \label{eq:5,2}
\end{eqnarray}
This equation corresponds to the field equation
$\delta S/\delta \varphi_i=0$ of the operator formalism.
The second term of \eqno(5,2) is a field-equation
violating term due to the use of the T$^*$-product.
One should never confuse it with the field-equation anomaly.
For example, in the conformal-gauge two-dimensional quantum gravity,
\eqno(4,14) is reproduced from \eqno(5,2) by setting $F=\,\tilde{\! b}{}^\lambda$
and $\varphi_i=\,\tilde{\! b}{}^\rho$. Likewise, if we set $F=\,\tilde{\! b}{}^\lambda$ and
$\varphi_i=h_\rho$ in \eqno(5,2), we obtain
\begin{eqnarray}
&&\taustar{\,\tilde{\! b}{}^\lambda(\,\tilde{\! b}{}^\rho-\skew3\widetilde{\tcalT}^\rho)}=0, \label{eq:5,3}
\end{eqnarray}
that is, we do not encounter the field-equation anomaly.
Instead, as is shown in the perturbative approach, \eqno(5,3) induces
the violation of the BRS invariance in the path-integral approach.
\par
Historically, the anomaly problem in the conformal-gauge two-dimensional
quantum gravity was discussed first by Fujikawa\cite{Fujikawa} in the
path-integral formalism.
His formulation is not, however, understandable in the framework stated
in \S4.
\par
First, he takes all {\it three\/} degrees of freedom of $g_{\mu\nu}$
as path-integration variables; nevertheless, each of his ghost fields
has only {\it two\/} degrees of freedom. The extra one degree of freedom
is the conformal one, denoted by $\rho$, is
{\it not allowed\/}\footnote{If one carries out the integration over
$\rho$ after introducing the
conformal ghosts, then it becomes impossible to work out his analysis.}
to be integrated (until the Liouville action is derived) in spite of the
fact that it is an independent path-integration variable.
\par
Second, by introducing tilde fields $\tilde\varphi{}_i=\rho^{n_i}\varphi_i$,
$n_i$ being a certain fractional number, he claims that the path-integral
measure becomes BRS invariant if it is expressed in terms of the tilde
fields. He then derives the Liouville action expressed in terms of
$\rho$ alone by calculating the variation of the path-integral measure
under the conformal transformation. That is, according to his theory,
the conformal anomaly is directly obtained {\it without passing
through the BRS anomaly\/} in contradiction to the consideration
presented in \S4.
\par
We are thus unable to reproduce his analysis in terms of the explicit
solution.
\section{FP-ghost number current anomaly}
As we emphasized previously,\cite{AN4} there is no FP-ghost number anomaly
in the conformal-gauge two-dimensional quantum gravity: The exact solution
is completely consistent with the FP-ghost number conservation.
The conservation of the FP-ghost number current $j_c{}^\mu$ is a simple
consequence of the fact that $c^\pm(x)$ and ${\skew3\bar c}\hspace*{1pt}^\pm(x)$ are independent
of $x^\mp$. This property is never violated at the representation level.
Nevertheless, many authors have claimed that the FP-ghost number current
has anomaly. The reasons for the occurrence of this belief are its
correspondence to the Riemann-Roch theorem and the field-equation-violating
property of the T$^*$-product.
\par
Fujikawa\cite{Fujikawa} was the first to claim the existence of the
FP-ghost number current anomaly. He derived it by making the FP-ghost
number transformation\footnote{This must be done in the original
tilde-variable expression. No anomaly is derived by the naive calculation
if this transformation is made in the equivalent expression having the
Liouville action explicitly.} in his path-integral formalism described
at the end of \S5. His result is written as
\begin{eqnarray}
&&\partial_\lambda\langle\!\langle\,j_c{}^\lambda\,\rangle\!\rangle
={3\over 4\pi}\langle\!\langle\, \partial^2 \log\rho \,\rangle\!\rangle,
\label{eq:6,1}
\end{eqnarray}
where we denote the path integration by
$\langle\!\langle\;\cdots\;\rangle\!\rangle$.
If the degrees of the reparametrization freedom is suppressed, one may
write $-\partial^2\log\rho=\sqrt{g}R$ (Euclidean metric is used).
Here we must note that Fujikawa's theory is formulated in the {\it flat\/}
background metric and that $\rho$ is the path-integration variable.
\par
Shortly later, Friedan, Martinec and Shenker,\cite{FMS} who formulated
conformal field theory, quoted \eqno(6,1) in the disguised form.
They consider a completely {\it curved\/} background metric $\hat{g}{}_{\mu\nu}$.
The right-hand side of their equation is const.$\sqrt{\hat{g}{}}\hat{R}$,
{\it a function of\/} $\hat{g}{}_{\mu\nu}$, which is nothing but the quantity
required by the Riemann-Roch theorem under the prerequisite of the
conformal covariance.
It may be certainly {\it analogous\/} to \eqno(6,1), but we cannot find any
{\it logical connection\/} between them.\footnote{According to Fujikawa
(private communication), the possible existence of the zero points of $\rho$
can take care of the effect of the topological number $\int d^2x\,\sqrt{\hat{g}{}}
\hat{R}$. We do not see how this idea can be formulated.}
\par
In the perturbative approach, the Friedan-Martinec-Shenker version of
\eqno(6,1) is interpreted, through the consideration based on the effective
action, as the matter that\cite{Dusedau,KR}
\begin{eqnarray}
&&J^\lambda{}_{\mu\nu}\equiv\taustar{j_c{}^\lambda
{\delta S\over \delta \hat{g}{}^{\mu\nu}}}|_{\hat{g}{}^{\mu\nu}=\eta^{\mu\nu}}
\label{eq:6,2}
\end{eqnarray}
has a nonvanishing nonlocal term, where $\hat{g}{}_{\mu\nu}$ is a background
metric introduced in such a way that the gauge-fixing plus FP-ghost
Lagrangian density becomes background covariant.
Note in \eqno(6,2) that the background metric is taken to be {\it flat\/}
in the Feynman-diagram calculation.
\par
In the conformal-gauge case, $J^\lambda{}_{\mu\nu}$ is essentially equal to
\begin{eqnarray}
&&\taustar{j_c{}^\pm(x_1)\skew3\widetilde{\calT}^\pm(x_2)}
=-12\partial_\pmD_{\hbox{\tiny F}}(x_1-x_2)\cdot\partial_\pm{}^2D_{\hbox{\tiny F}}(x_1-x_2).
\label{eq:6,3}
\end{eqnarray}
It is in this sense that the FP-ghost number current anomaly is claimed
to be obtained in the perturbative approach.
It should be noted, however, that {\it if the T$^*$-product is not taken,
that is, if $D_{\hbox{\tiny F}}$ is replaced by $D^{\hbox{\tiny (+)}}$, \eqno(6,3) becomes consistent with\/}
$\partial_\lambda j_c{}^\lambda=0$.
\par
D\"usedau\cite{Dusedau} found that $J^\lambda{}_{\mu\nu}$ has no nonlocal
term in the de Donder-gauge case. Kraemmer and Rebhan\cite{KR} discussed
the gauge dependence of $J^\lambda{}_{\mu\nu}$ and claimed that the gauge
independence can be recovered if one adds a contribution from the
``Lagrange-multiplier (or B-field) current''.\footnote{Although we cannot
regard their proof as adequate, their claim itself can be verified by
explicit calculation.\cite{AN9}} But the relation between this fact and
the Riemann-Roch theorem was not discussed.
\par
Recently, Takahashi\cite{Takahashi} has reconsidered D\"usedau's analysis
from his perturbative approach described in \S4.
In discussing $j_c{}^\lambda$, he regards the quantum gravitational field
as the background metric, just as Fujikawa did. Rederiving D\"usedau's
result, he asserts that the vanishing of the FP-ghost number current
anomaly can be explained by the existence of the FP-ghost conjugation
invariance of the de Donder gauge two-dimensional quantum gravity.
One should note, however, that, as pointed out in \S2, the {\it FP-ghost
conjugation invariance exists also in the conformal-gauge case}.
Therefore, his standpoint would imply the absence of the FP-ghost number
current anomaly also in the conformal-gauge case.
\par
Finally, we note that the nonexistence of the FP-ghost number current
anomaly can be shown even if the gauge-fixing background metric is
a nonflat one given by $\hat\rho(x)\eta_{\mu\nu}$, where
$\hat\rho{}^{-1}$ {\it is assumed to exist}.
In this case, \eqno(2,6) is replaced by
\begin{eqnarray}
&&\tg^{\mu\nu}=(\eta^{\mu\nu}+\hat\rho^{-1} h^{\mu\nu})
(1-\hat\rho{}^{-2}\det{h^{\sigma\tau}})^{-1/2}.
\end{eqnarray}
The Lagrangian density of this case is obtained from \eqno(2,8) by
simply replacing $h_\lambda$ by $\hat\rho^{-1}h_\lambda$.
Since ${\mbf\delta_*}\,(\hat\rho^{-1}h_\lambda)=\hat\rho^{-1}{\mbf\delta_*}\,(h_\lambda)$,
we can absorb the factor $\hat\rho^{-1}$ into $\,\tilde{\! b}{}^\lambda$ and
${\skew3\bar c}\hspace*{1pt}^\lambda$ by redefining them. Thus the ghost part of
the Lagrangian density of the nonflat case becomes completely the same
as that of the flat case. Thus nothing new can happen about the
FP-ghost number current.
\section{Discussion}
Nowadays, the path-integral approach and the counter-term business have
become so fashionable that many physicists preclude the consideration
based on other approaches from the outset. Certainly, the path-integral
approach is convenient and successful in gauge theories, but we wish to
emphasize that the same is not necessarily true in quantum gravity.
\par
The path-integral formalism directly deals with the solution at the
representation level. Accordingly, if one adopts the path-integral
approach, one can no longer perceive what happens in the transition from
the operator level to the representation level. Indeed, one cannot
describe the existence of the field-equation anomaly in the path-integral
approach.
\par
The quantities describable by the path-integral formalism are those
which can be written in terms of the T$^*$-product. The T$^*$-product is
certainly a very convenient notion because we need not take care of the
ordering problem even for the timelikely separated field operators.
On the other hand, as emphasized in the present paper, the T$^*$-product
has a demerit of violating the field equations explicitly.
As demonstrated in the present paper, this fact induces unpleasant
complications and misleading expressions.
Furthermore, since the T$^*$-product contains $\theta$-functions, the
Green's function is more singular than the corresponding Wightman functions,
that is, some singularities found in the perturbative or path-integral
approach may be superficial. When this fact is combined with the
counter-term business, one is led to introducing counter terms which
are purely of the T$^*$-product origin. In the present paper, we have
demonstrated that ``anomalies'' also can be of the T$^*$-product origin.
\par
We hope that more physicists reinvestigate the anomaly problem in
quantum gravity without adhering to the path-integral approach.
\section*{Acknowledgements}
One of the present authors (N. N.) would like to express his sincere
thanks to Professor K. Fujikawa and Dr. H. Kanno for the discussions
concerning the FP-ghost number current anomaly.
\ifundefined{abstsize}
\newpage
\fi
|
\section{Introduction}
Some Calabi-Yau three-folds(=CY 3-folds) with base, \relax {\bf \rm C}{\rm I\kern-.18em P}$^1(1,s)$ and fiber,
\newline
K3$= \relax {\bf \rm C}{\rm I\kern-.18em P}^3(u_1, u_2, u_3, u_4)[d]$
are represented in hypersurface in weighted projective
4-space, \relax {\bf \rm C}{\rm I\kern-.18em P}$^4$.
\begin{equation}
\CY3 {\rm -fold} =
\relax {\bf \rm C}{\rm I\kern-.18em P}^4(u_1, s u_1, (s+1)u_2, (s+1)u_3, (s+1)u_4))[(s+1)d],
\end {equation}
\noindent
with $ d= \sum_{i=1}^4 u_i$.
\par
The type IIA string on
CY3-folds which have K3 fiber and T$^2$ fiber with at least one
section is dual to the heterotic string on K3 $\times$ T$^2$
as pointed out in~\cite{vafa1,vafa2}.
Much research has been explored on CY3-folds with
$s=1$ and $u_1$=1 cases and F$_0$ based case~\cite
{vafa1,vafa2,aldazabal,bershadsky,candelas1,candelas2,candelas3}.
$s=1$ and $u_1$=1 cases are given by the
extremal transitions
or by the conifold transitions
from F$_0$ based CY3-folds \cite{candelas1,candelas2,candelas3}.
Thus far, CY3-folds which have been studied
are constructed by using dual polyhedra.
Some of them may be related to the hypersurface representations of eq. (1)
with $s \geq 2$. However,
identification between them
has some ambiguities and is complicated~\cite
{candelas4,candelas0,skarke,avram}.
Therefore, the relation between these results and the
duality between the type IIA string on CY3-folds with $s \geq 2$
to heterotic string on K3 $\times$ T$^2$
is not clear yet.
There have been three types of web sequences of
heterotic - type IIA string duality
from the terminal CY3-fold in A series
\cite{aldazabal,candelas1,candelas2}.
(iii) in \cite{candelas2} may be the subset of (i)$^\dag$ in
\cite{candelas3},
though the properties of dual polyhedron
are slightly different
\footnote{ The difference between
dual polyhedron of (i)$^\dag $
and (iii) is as follows \cite{candelas1,candelas2,candelas3}.
The dual polyhedra of case (i)$^\dag$ have the modified
dual polyhedra of K3 fiber.
For the case (iii) in A series, the dual polyhedron of
K3 fiber is not modified.
The highest point in the
additional points is represented by the weights of the
K3 fiber of the terminal A series in this base.
One point such as $(0,\ast,\ast,\ast)$ is also
represented by the part of the weight of the terminal K3 fiber.
The following element in SL(4,{\bf Z}) can
transform these polyhedra
into the dual polyhedron given by
the ref. of \cite{candelas2}.
\[
\left(
\begin{array}{cccc}
1 & 0&1&2 \\
0&1& 2& 3 \\
0&0 & 1 &2\\
0&0 & -1& -1\\
\end{array}
\right).
\]
In the base of \cite{candelas2},
the additional points make a line with $x_4=-1$. }.
(i)$^\dag$ means the modified (i) in
\cite {candelas1} with extra tensor multiplets.
\begin{eqnarray}
&{\rm (i)}&
G_2^0=I \mid_{G_1^0, {n_T}^0} ~{\rm with}~ n^0 = \{ 0,1, 2 \}
\rightarrow G_2^0=\hat G_2 \mid_ {G_1^0,{n_T}^0}~ {\rm with}~ n^0 \geq 3
\rightarrow \cdots,
\nonumber \\
& \big\{ {\rm (i)}^\dag &
(G_2^0, {n_T}^0) \mid_{G_1^0 }
\rightarrow (G_2, {n_T}) \mid_ {G_1=G_1^0}
\rightarrow \cdots ,\big\}
\nonumber \\
&{\rm (ii)}& G_1^0=I \mid_{G_2^0,n_T^0}~{\rm with}~ n^0=j \rightarrow
{G_1^0}=\hat G_1 \mid_ {G_2^0, n_T^0}~ {\rm with}~ n^0=j \rightarrow \cdots,
\nonumber \\
& \big\{ {\rm (ii )}^\dag & (G_1^0,n_T^0)\mid_{G_2^0} \rightarrow
( {G_1}, n_T)\mid_ {G_2=G_2^0} \rightarrow \cdots, \big\}
\nonumber \\
&{\rm (iii)}&
(G_2^0=I, n_T^0 )\mid _{G_1^0=I}~{\rm with}~ n^0=0 \rightarrow (G_2,n_T)
\mid_{G_1=I} ~
\rightarrow \cdots,
\end{eqnarray}
\noindent
where $\hat G_1$ and $\hat G_2$ are non-Abelian gauge
symmetries and $ 12 \geq j \geq 0$, see Fig. 1 and Fig. 2.
In this paper, we investigate the possibility of
following sequence where we picked up CY3-folds and Hodge numbers
from the list by Hosono et al. \cite{yau}.
\begin{equation}
{\rm (iv)} ~
(G_2^0=I, s=1,n_T^0)\mid_{ G_1^0=I} ~{\rm with}~ n^0=2
\rightarrow (G_2,s=2,n_T)\mid _{G_1=I} \rightarrow \cdots
\end{equation}
\begin{eqnarray}
&G_1^0~ ({\rm or}~G_1)&:{\rm the ~gauge ~symmetry~with ~charged~ matter,~whose~information}
\nonumber \\
&~&{\rm ~may~come~ from ~ the~ bottom~ of~dual~polyhedron,}
~\nabla ~
\nonumber \\
&~&{\rm of~ appropriate~ K3~ fiber~ for~ the~ series~ (ii)~(or~(ii}^\dag))
~\cite{candelas1},
\nonumber \\
&G_2^0~({\rm or}~G_2) &:{\rm the~terminal ~gauge ~symmetry~with~ no~ charged~ matter, ~whose}
\nonumber \\
& &{\rm ~information~may~come~ from ~ the~ top~ of~dual~polyhedron,}
\nonumber \\
&& ~\nabla~ {\rm of~ appropriate~K3~ fiber~ for~ the~ series~ (i)~( or~ (i
}^\dag)) ~\cite{candelas1,skarke}.
\end{eqnarray}
\noindent
For A series, the non-Abelian instanton numbers
of the vector bundles on K3 are denoted as
$(k_1^0,k_2^0);~ k_1^0+k_2^0=24$~
\footnote{
Let $m_i^0=m_i$, $m_{iA}^0=m_{iA}$ and $m_{iB}^0 =m_{iB}~ (i=1,2)$ be Abelian instanton numbers.
For B series, non-Abelian instanton numbers and Abelian instanton numbers
are denoted as
$(k_1^0,m_1^0:k_2^0,m_2^0);~ k_1^0+k_2^0=18,~ m_1^0=m_2^0=3$.
For C series, they are denoted as
$(k_1^0,m_{1A}^0,m_{2A}^0:k_2^0,m_{1B}^0,m_{2B}^0),
~ k_1^0+k_2^0=14,~ m_{1A}^0=m_{1B}^0=3,~
m_{2A}^0=m_{2B}^0=2$ .}.
\noindent
The integer of $n^0$ is defined by $k_1^0=12+n^0$ and
~$k_2^0=12-n^0$~
\footnote{
The number of non-Abelian instantons should
be greater than three except zero.
The terminal group should change to avoid
this situation, which causes a correction or a modification
\cite{aldazabal,candelas2,candelas3}.
For C series, the non-Abelian and Abelian instanton numbers
are modified as follows \cite{aldazabal}.
In $n^0=5$ case, $ (k_1^0,m_{1A}^0,m_{1B}^0~;~
k_2^0,m_{2A}^0,m_{2B}^0)=(12,3,2;0,3,3)$.
In $n^0=6$ case, $(k_1^0,m_{1A}^0,m_{1B}^0~;~
k_2^0,m_{2A}^0,m_{2B}^0)=(13,3,2;0,3,3)$.}.
The suffix 0 denotes terminal case with
$ n_T=1 $ in each series, where
``terminal case'' means $G_1^0=I$ or $G_1=I$.
\noindent
When $\Delta n_T$
E$_8$ small instantons shrink, in
the terminal gauge symmetry side, the
non-Abelian instanton numbers should be modified as follows
\cite{aldazabal,candelas3}:
$(k_1,k_2)$,~$ k_1+k_2+\Delta n_T=24$,~$k_1=k_1^0 = 12$,~
$k_2=k_2^0-\Delta n_T=12-
\Delta n_T$
\footnote
{Similarly to A, for B and C, the modifications may follow
\newline
\noindent
B-chain~:
$k_1 + m_1=k_1^0 +m_1^0 = 12$,~
$k_2+m_2 =12 -\Delta n_T $,
\newline
C-chain ~: $ k_1+m_{1A}+m_{1B} = k_1^0+m_{1A}^0+m_{1B}^0 =
12 $,~$k_2+m_{2A}+m_{2B} =12 -\Delta n_T $,
\newline
which we will discuss in this article.}.
The additional $\Delta n_T$ numbers of tensor multiplets
are created and 29 $\Delta n_T$ appears in the anomaly
cancellation condition
\footnote{
29 is the dual coxetor number of E$_8 - $one \cite{aldazabal,candelas3}.
The subtraction of one means the freedom of fixing the
place where a tensor multiplet
is created.
The dual coxetor number of E$_7 -$one = 17 works for B-chain case
when a tensor multiplet is created and that of E$_6 -$one =11 for
the C-chain case \cite{aldazabal,candelas3}.}.
$n_T=n_T^0+\Delta n_T$ with $ n_T^0=1$ is
the number of the tensor multiplets in D=6 N=1 compactification
\footnote{In D=4 and N=2 case, these tensor multiplets become
vector multiplets. }.
\vspace{8pt}
The puzzles which we would like to consider
are as follows:
\newline
I:~Why (iii) and (iv) have the same Hodge numbers for A series ?
Do (iv) satisfy the duality ? Though K3 fibers of CY3-folds in (iv)
are different from those in (iii), is there
any gauge enhancement for (iv) ?
\newline
II:~Is it possible to obtain
B or C versions of (iii) which satisfy the duality
by means of conifold transitions ?
Already, the method to
construct dual polyhedron was given by \cite{candelas2,candelas3}
~about this question.
They pointed out that their Hodge numbers are obtained by shrinking
of tensor multiplets \cite{candelas3}.
\newline
III:~Is it possible to obtain
B or C versions of (iv) which satisfy the duality by means of conifold
transitions ?
\newline
IV:~ Do the Hodge numbers in B or C series of (iii)
coincide with those of (iv)?
\newline
V:~ Are there any
hypersurface representing CY3-folds with $s \geq 2$
which satisfy the duality from $n^0 \geq 4$ ?
\newline
In section 2, we investigate prob. I.
We also discuss prob. III by using the results
of \cite{aldazabal}
and \cite{candelas3} in section 3.
\newline
\section{The duality in K3 fibered Calabi-Yau 3-fold}
\par
~~~~~ We treat the extension from A-chain $n^0=2$ terminal case.
$\relax{\rm CY}_3=\relax {\bf \rm C}{\rm I\kern-.18em P}^4(1,s,(s+1)(1,4,6))[12(s+1)]$ with
$\relax{\rm K3}=\relax {\bf \rm C}{\rm I\kern-.18em P}^3(1,1,4,6)[12]$ fiber
have the same Hodge numbers as those examples
given by \cite{candelas2,yau}.
To see why they coincide, we
compare dual polyhedron of CY 3-folds.
They are not SL(4,{\bf Z}) equivalent.
We represent them in the base where
one of vertices of
dual polyhedron in the bottom
denotes the part of its weight
for the hypersurface representation. It is
~$(-s,-s-1,-4(s+1),-6(s+1))$ in this case.
The base manifold under the elliptic fibration is F$_2$ for $s=1$ case.
F$_i$ denotes Hirzebruch surface.
The dual polyhedron of
F$_i$ has three vertices~:~$ \{ \vec v_1,\vec v_2,
-\vec v_1 -i\vec v_2 \}$.
An example of integral points in the dual polyhedron of F$_i$ is given by
$\{\vec v_1=(1,0) ,(0,-1), \vec v_2=(0,1),(0,0),(-1,-i)\}$.
The Hodge numbers
and the dualities in case (iii) are derived
by investigating the extremal transition of the dual polyhedron
of F$_0$ based CY3-fold in \cite{candelas2}.
By using dual polyhedra, we can find a
fibrations and base manifolds in some cases
\cite{candelas1,candelas3,skarke,avram}.
The upper suffix in $\nabla$
denotes the dimension of the lattice of a polyhedron.
$\{(x_1,x_2,x_3,x_4)\}
~\subset {}^4\nabla $~ forms the
dual polyhedron of CY3-fold.
$\{(x_1,x_2)\}~\subset {}^2\nabla~ \subset {}^4 \nabla $ represents
the dual polyhedron of the base under the
elliptic fibration of CY3-folds.
In this paper, they are the blown up Hirzebruch surfaces.
$\{(x_2,x_3,x_4)\} \mid_{x_1=0}
~\subset {}^3\nabla ~ \subset {}^4\nabla $
represents
the dual polyhedron
of K3 fiber of CY3-fold.
$ \{ (x_3,x_4) \}
\mid_{x_1=x_2=0} ~\subset {}^2\nabla $~$~ \subset {}^4\nabla~$ are
the dual polyhedron of common
elliptic fiber of CY3-folds and K3fiber. However,
this is not a sufficient condition of having elliptic fibration.
The result of fibrations in CY3-folds is given in table 1.
\begin{table}
\[
\begin{array}
{|l|l|l|l|l|l|}
\hline
\multicolumn{3}{|c|}
{\rm CY3-folds ~of~ \cite{candelas2}} &
\multicolumn{3}{|c|}
{\rm CY3-folds ~ of~ hyper ~ surface~ rep.}\\
\hline
\Delta n _T & {\rm the~ base~ under~}T^2 & {\rm K3~ fiber} & s
& {\rm the~ base~ under~}T^2 & {\rm K3~ fiber}
\\ \hline
0 & {\rm F}_0 & \relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,1,4,6)[12]&1&{\rm F} _2&CP ^3 (1,1,4,6)[12] \\ \hline
2 & {\rm F} _2~ {\rm blown ~ up} & \relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,1,4,6)[12]&2&{\rm F} _2~
{\rm blown~up}&CP ^3 (1,1,4,6)[12] \\ \hline
3 & {\rm F} _3 ~{\rm blown~ up}
& \relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,2,6,9)[18]&3&{\rm F} _2~
{\rm blown~ up}&\relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,1,4,6)[12] \\ \hline
4 & {\rm F} _4~{\rm blown ~up} & \relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,2,6,9)[18]&4& {\rm F} _2~ {\rm blown~ up}&\relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,1,4,6)[12] \\ \hline
6 & {\rm F} _6~{\rm blown ~up}
& \relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,3,8,12)[24]&6& {\rm F} _2~{\rm blown ~up}&CP ^3 (1,1,4,6)[12] \\ \hline
8 & {\rm F} _8~{\rm blown ~up}
& \relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,4,10,15)[30]&8&{\rm F }_2~{\rm blown~ up}&
\relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,1,4,6)[12] \\ \hline
12 & {\rm F} _{12}~{\rm blown ~up}
& \relax {\bf \rm C}{\rm I\kern-.18em P}^3(1,5,12,18)[36]&12&{\rm F} _2~{\rm blown~ up} &
\relax {\bf \rm C}{\rm I\kern-.18em P} ^3 (1,1,4,6)[12] \\ \hline
\end{array}
\]
\caption{ The kinds of fibrations of CY3-folds}
\normalmarginpar{N.B.~;~ The dual polyhedron of \relax {\bf \rm C}{\rm I\kern-.18em P}$^3$(1,5,12,18)[36]
coincides with that of \relax {\bf \rm C}{\rm I\kern-.18em P}$^3$(1,6,14,21)[42].}
\end{table}
In $s \geq2$ case, the additional integral points to F$_2$ are
$\{(-i,-i-1) ~ {\rm with}~
\newline
2 \leq i \leq s,~(-1,-1) \} $ within the base.
Thus, for $s \geq2$ case, the base manifolds remain blown up F$_2$
since ${ s+1 \over s} \leq 2$.
We can see the Hodge number of the base under the elliptic
fibration and $n_T$ from dual polyhedron,
$n_T+1 = $ $h^{1,1}({\rm F}_2~{\rm blown ~up})=
d_1-2d_0=s+2 ~(s \geq 2),$
where $d_i$ is the number of
i-dimensional cones \cite{fulton}.
$n_T$ coincides with case (iii) \cite{candelas2}.
On the other side,
K3 fibrations of case (iii)
change so that the base manifolds under elliptic fibrations
also alter to F$_i$ blown up $(i \geq 2)$ with
including F$_0$ and F$_2$.
In any case, the bases under elliptic fibrations in case (iv) are
birationally equivalent to those in case (iii),
which leads to the following identification.
(iii) and (iv) are connected by the extremal
transitions within each sequences. $ \Delta n_T=0$ cases are
deformed to each other
by the change of base manifolds, i.e.,
F$_0 \leftrightarrow$ F$_2$ by using
a non-polynomial freedom of deformation \cite{candelas2,gross}.
Both CY 3-folds with $\Delta n_T=0 $ are the same manifold
with double K3 fibrations \cite{gross}.
\begin{eqnarray}
\hat \nabla_{s=12, \Delta n_T=12}
\supset
\cdots
\hat \nabla_{s=3,\Delta n_T=3}
\supset
\hat \nabla_{s=2,\Delta n_T=2}
\supset
&\hat \nabla_{s=1,\Delta n_T=0}&
\nonumber \\
&\updownarrow &
\nonumber \\
\nabla_{\Delta n_T=12}
\supset
\cdots
\supset
\nabla_{\Delta n_T=3}
\supset
\nabla_{\Delta n_T=2}
\supset
&\nabla_{\Delta n_T=0}&.
\end{eqnarray}
\noindent
For $\Delta n_T > 0$,
blown up F$_2$ based CY3-folds
in case (iv)
also can be deformed to
those of case (iii) in \cite{candelas2} by using a non-polynomial
freedom of deformation.
$
\hat \nabla_{s=i, \Delta n_T =i} \leftrightarrow
\nabla_{\Delta n_T=i},
$
\noindent
where
$\hat \nabla$ denotes dual polyhedron of
case (iv) and $\nabla$
denotes those of case (iii)
\cite{candelas2}.
The CY 3-folds
constructed by
\cite{candelas2} and those with $s$
are the same manifolds
with double K3 fibrations \footnote{They are represented by the
same dual polyhedra in \cite{candelas3}.}.
Each CY 3-fold with $\Delta n_T=i$ can be
represented as a K3 fibration in two inequivalent ways as table 1.
The CY3-fold with $s=1$ is
the terminal case of A-chains with $n^0=2$ in
the duality web \cite{aldazabal}.
If duality exists, then
their Hodge numbers must
satisfy the following conditions which
comes from D=6 and D=4 theories as the
anomaly cancellation \cite{vafa1,candelas1,candelas2}.
\begin{eqnarray}
h_{2,1}\mid _{ G_2~ {\rm in~ (iv)} }&=&h_{2,1}^0
\mid_{G_2^0=I~{\rm in ~(i)}}
-( a-b\Delta n_T) + {\rm dim.} G_2-29 \Delta n_T,
\\ \nonumber
h_{1,1} \mid_ {G_2 ~{\rm in~(iv)}} &= &
h_{1,1}^0 \mid_{G_2^0=I~{\rm in ~(i)}} + {\rm rank}G_2
+ \Delta n_T,
\\ \nonumber
h_{2,1}^0
\mid_{G_2^0=I~{\rm in ~(i)}} &=&243,
~h_{1,1}^0 \mid_{G_2^0=I~{\rm ~in~ (i)}}=3.
\end {eqnarray}
\noindent
where $G_1^0=G_1 =I$ up to U(1) in these cases.
\noindent
This duality between heterotic string and Type IIA string is
summarized in the
table 4, which coincides with case (iii)
\cite{candelas2}\footnote{ This may imply heterotic -
heterotic string duality.}.
For CY3-folds side, $G_1$ and $G_2$ symmetries
are due to the quotient singularities
of K3 fiber in the first column of table 1
rather than the quotient singularities of CY 3-folds with s in
table 2
\footnote{
We compared some superpotentials of type IIB side in case (iii) and
those in case (iv) and examined the possibility to derive the
terminal gauge symmetry by the method of \cite{lerche}.
We will give dual polyhedra of case (iv) in the
next paper to report about superpotentials
more precisely.} .
\begin{table}[h]
\[
\begin{array}{|l|l|l|l|l|l|l|l|}
\hline
s & s=1 & s=2 & s=3 & s=4 & s=6
& s=8 & s=12
\\ \hline
{\rm quot. sing.} & A_1 &A_1A_2 & A_1^2 A_3
& A_1^4 A_4 & A_1^3 A_6 & A_1^4 A_8 & A_1^6 A_{12}
\\ \hline
\end{array}
\]
\caption{The quotient singularities in K3=\relax {\bf \rm C}{\rm I\kern-.18em P}$^3$(1,1,4,6)[12] fibered CY3-fold }
\end{table}
For K3$ \times$ T$^2$ side,
they are the subgroups of the E$_8 \times $ E$_8$ group
for case (iii) \footnote{
They are perturbative gauge symmetries for case (iii)
and may be non-perturbative ones for case (iv)
as $ \Delta n_T=0 $ case \cite{gross}.}.
$a+ bn $ is calculated by the index theorem and denotes
the number of $G_1$ (or $G_2$) charged hyper multiplet fields, which is
summarized in table 3~\cite{bershadsky,aldazabal,candelas1,candelas3}.
$n=n^0$ for $ G_1^0= \hat G_1$ case.
The number of the charged hyper multiplets of $G_2$
vanishes. We can see this by substituting $n=-n^0$ for $G_2^0= \hat G_2$ case
and $n=-\Delta n_T$ for $G_2=\hat G_2$ for A series ~\cite{candelas2,
candelas3}.
\begin{table}[h]
\[
\begin{array}
{|l|l|l|l|}\hline
G_1 & {\rm A~ series} &{\rm B~ series} &{\rm C~ series}
\\ \hline
A_1 & 12n+32 &6k_1-39 & 4k_1-17
\\ \hline
A_2 & 18n+54 & 10k_1-64 & 6k_1-26
\\ \hline
A_3 & 22n+76 &12k_1-75 & 8k_1-33
\\ \hline
A_4 & 25n+100 &14k_1-84 & 9k_1-35
\\ \hline
D_4 & 24n+96 & &
\\ \hline
D_5 & 26n+124 &15k_1-90 & 10k_1-39
\\ \hline
E_6 & 27n+162 &16k_1-94 &
\\ \hline
E_7 & 28n+224 & & \\ \hline
\end{array}
\]
\caption{ The review of the number of the charged hyper multiplets.}
\end{table}
\begin{table}[h]
\[
\begin{array}{|l|l|l|l|l|l|l|l|l|l|}\hline
s & U(1)^4 \times G_2 & h^{1,1}& h^{1,2}& k_1& k_2
& n_ T^0&\Delta n_T& n_T &n^0 \\ \hline
s=1 & U(1)^4 \times (G_2^0=I) &3 &243 & 12 & 12 & 1&0&1 &
2 \\ \hline
s=2 & U(1)^4 \times I &5 &185 & 12 & 12-2& 1&2&3& \\ \hline
s=3 & U(1)^4 \times A_2 &8 &164 & 12 &12-3 & 1&3&4 &\\ \hline
s=4 & U(1)^4 \times D_4 &11 &155 & 12 & 12-4 & 1&4&5 &\\ \hline
s=6 & U(1)^4 \times E_6 &15 &147 & 12 & 12-6 & 1&6&7 &\\ \hline
s=8 & U(1)^4 \times E_7 &18 &144 & 12 & 12-8 & 1&8&9 &\\ \hline
s=12 & U(1)^4 \times E_8 &23 &143 & 12 & 12-12 & 1&12&13 &\\ \hline
\end{array}
\]
\caption{The duality of (iv) about
K3=\relax {\bf \rm C}{\rm I\kern-.18em P}$^3$(1,1,4,6)[12] fibered CY3-fold }
\end{table}
\section{Discussion and Conclusion}
\par In this article, we have studied the property of CY3-folds
with $ s\geq 2$ case from A series.
They are the same CY 3-folds as those constructed in \cite{candelas2},
which have double K3 fibrations.
\newpage
For the puzzle III, the conclusion is that
the hypersurface representations with
$s \geq 2$ cases
from B series and C series in $n^0=2$ satisfy
the duality of terminal B and C series with extra tensor multiplets
\footnote{
For B and C sequences, the modification of eq. (6) is necessary.
The change in the Hodge numbers, $h_{2,1}$ according to the change
of the terminal group is not only the difference of
dim. $G_2^0$~( or dim. $G_2$)~\cite{aldazabal,candelas3}.
Thus we compare the Hodge numbers of CY3-folds which have
the same terminal gauge symmetry.}.
For example, B and C series in s=2, their
Hodge numbers can be interpreted as those with shrinking of
two instantons from
$n^0=2$ case as the ref. of \cite{candelas3}.
$s \geq 3$
cases are also explained by shrinking $\Delta n_T$ instantons
from the terminal case with the same $\hat G_2$ symmetry.
\begin{eqnarray}
&\Delta h_{2,1}&= -h_{2,1}^0 \mid_{G_2^0=\hat G_2~{ \rm terminal~in~(i)}}
+h_{2,1} \mid_{G_2=\hat G_2~{ \rm terminal~in~(iv)}},
\nonumber \\
&\Delta h_{1,1}&= -h_{1,1}^0 \mid_{G_2^0=\hat G_2~{ \rm terminal~in~(i)}}
+h_{1,1} \mid_{G_2=\hat G_2~{ \rm terminal~in~(iv)}},
\nonumber \\
&{\rm B~ series}&~ : -\Delta h_{2,1}=17 \Delta h_{1,1},
~{\rm C~ series} : -\Delta h_{2,1}=11 \Delta h_{1,1}.
\end{eqnarray}
\noindent
This situation is quite similar to that of A series.
They are interpreted as the dualities obtained by
unhiggsing of $U(1)$ and $U(1)^2$
from (iv) in A series.
For puzzle IV,
Candelas, Perevalov and Rajesh seemed to
construct B and C versions of (iii) and derived
the Hodge numbers already, though they do not write them
explicitly \cite{candelas3}.
According to the description,
Hodge numbers of them are the same as table 5 and table 7.
\begin{table}
\[
\begin{array}{|l|l|l|l|l|l|l|l|l|l|}\hline
s & U(1)^4 \times U(1) \times G_2 & h^{1,1}&h^{1,2} &
k_1&k_2&{n_T}^0&\Delta n_T& n_T &n^0\\
\hline
s=1 & U(1)^4 \times U(1) \times (G_2^0=I) & 4 & 148 & 9 & 9 &1
& 0 & 1& 2\\ \hline
s=2 & U(1)^4 \times U(1) \times I & 6 & 114
& 9 & 9-2
& 1&2&3 & \\ \hline
s=3&U(1)^4 \times U(1)\times A_2 & 9 & 101 & 9 & 9-3
& 1&3&4 &\\ \hline
s=4&U(1)^4 \times U(1) \times D_4 & 12 & 96
& 9 & 9 -4
& 1&4&5& \\ \hline
s=6&U(1)^4 \times U(1)\times E_6 & 16 & 92
& 9 & 9 -6
& 1&6&7& \\ \hline
\end{array}
\]
\caption{ The duality of (iv) about K3=\relax {\bf \rm C}{\rm I\kern-.18em P}$^3$(1,1,2,4)[10]
fiberd CY3-fold }
\end{table}
\begin{table}
\[
\begin{array}{|l|l|l|l|l|l|l|l|l|}\hline
n^0 & U(1)^4 \times U(1) \times G_2^0 & h^{1,1}&h^{1,2} &
k_1&k_2&{n_T}^0&\Delta n_T& n_T \\
\hline
n^0=0&U(1)^4\times U(1) \times I & 4 & 148 &9 & 9 &1
& 0&1 \\ \hline
n^0=3&U(1)^4 \times U(1) \times A_2 & 6 & 152 & 9+3 & 9-3
& 1&0&1 \\ \hline
n^0=4&U(1)^4 \times U(1)\times D_4 & 8 & 164
& 9+4 & 9 -4
& 1&0 &1 \\ \hline
n^0=6&U(1)^4 \times U(1) \times E_6 & 10 & 194
& 9+6 & 9 -6
& 1&0&1 \\ \hline
\end{array}
\]
\caption{ The duality of (i) about terminal CY3-folds in B series }
\end{table}
For the puzzle V, we examine K3 fiberd CY3-folds
whose s=1 case is
the terminal case in the A-chains with $n^0=4$.
Their K3 fiber is
\newline
K3=\relax {\bf \rm C}{\rm I\kern-.18em P}$^3$(1,2,6,9)[18].
The dual polyhedron of $s=2$ case has
following additional points
$(-2, -6,-18, -27)$, $(-1,-3,-10,-15)$ and $(-1, -3,-9,-14)$, which imply
$\Delta n_T$=1. $ h_{1,1} \mid _{s=2}-h_{1,1}^0 \mid _{s=1}=3$.
The terminal group should change
from $D_4$ to $G_2$ with rank 6. We can not find
$G_2$ which satisfies the extension of eq. (6).
It seems that it needs another idea to see the
duality in this case.
\newline
{\bf Acknowledgement}
~~~~~We gratefully acknowledge fruitful conversations with
M. Kobayashi, S. Hosono and N. Sakai. We would like to thank
K. Mohri mostly for useful discussions.
\begin{table}
\[
\begin{array}
{|l|l|l|l|l|l|l|l|l|l|} \hline
s& U(1) \times U(1)^2
\times G_2 & h^{1,1}& h^{1,2}&k_1&k_2&{ n_T }^0& \Delta n_T & n_T
&n^0
\\ \hline
s=1 & U(1)^4\times U(1)^2 \times (G_2^0=I) &5 & 101&7&7&1&0 &1&2\\ \hline
s=2 & U(1)^4 \times U(1)^2 \times I & 7& 79& 7& 7-2 &1&2&3 & \\ \hline
s=3 & U(1)^4 \times U(1)^2 \times A_2 & 10 &70& 7 & 7-3 &1&3&4 & \\ \hline
s=4 & U(1)^4 \times U(1)^2 \times D_5 & 13 &67& 7 & 7-4 &1&4&5 & \\ \hline
s=5 & U(1)^4 \times U(1)^2\times E_6 & 17 &65& 7 & 0 &1&5&6 & \\ \hline
s=6 & U(1)^4 \times U(1)^2\times E_6 & 17 &65& 7 & 0 &1&6&7 & \\ \hline
\end{array}
\]
\caption{The duality of (iv) about K3=\relax {\bf \rm C}{\rm I\kern-.18em P}$^3$(1,1,2,2)[6] fiberd CY3-fold
}
\end{table}
\begin{table}
\[
\begin{array}
{|l|l|l|l|l|l|l|l|l|} \hline
n^0& U(1) \times U(1)^2
\times G_2^0 & h^{1,1}& h^{1,2}&k_1&k_2&{ n_T }^0& \Delta n_T & n_T
\\ \hline
n^0=0 & U(1)^4\times U(1)^2\times I &5 & 101&7&7&1&0 &1\\ \hline
n^0=3 & U(1)^4 \times U(1)^2\times A_2 & 7 &103& 7+3& 7-3 &1&0&1 \\ \hline
n^0=4 & U(1)^4 \times U(1)^2 \times D_5 & 10 &110& 7+4 & 7-4 &1&0&1 \\ \hline
n^0=5 & U(1)^4 \times U(1)^2 \times E_6 & 12 &120& 7+5 & 0 &1&0&1 \\ \hline
n^0=6 & U(1)^4 \times U(1)^2 \times E_6 & 11 &131& 7+6 & 0 &1&0&1 \\ \hline
\end{array}
\]
\caption{The duality of (i) about terminal CY3-folds in C series
}
\end{table}
|
\section{Introduction}
A recent great development of nonperturbative aspects of
superstring theory has begun with following two events.
One is a proposal of M-theory \cite{Witten1}, which is a hypothetical
theory describing strongly coupled region of type IIA superstring
theory and reduces to eleven-dimensional supergravity in the low
energy limit.
The relation among parameters of string theory (the string length
$l_s$ and the string coupling $g_s$) and of eleven-dimensional
supergravity (the Planck length $l_{11}$ and
the radius of compactified the 11-th direction $R_{11}$) is given by
\begin{equation}
R_{11}=g_sl_s=g_s^{2/3}l_{11}.
\end{equation}
M-theory gives a systematic view point
for string theory --- in considering the M-theory,
all known perturbative
string theories are unified via various connections from duality.
The other is a discovery of D-branes \cite{Polchinski}, which are
solitonic solutions of string theory and can be constructed by using
conformal field theory in a well-defined manner. Thus they are
powerful tools for investigating nonperturbative phenomena in
string theory.
Further, it was followed by a new interpretation of
dimensional reductions of ten-dimensional supersymmetric
Yang-Mills (SYM) theory as low energy effective theories for
dynamics of the D-branes \cite{Witten2}.
Those movements culminated in the conjecture
by Banks, Fischler, Shenker and Susskind (BFSS) \cite{BFSS},
that ten-dimensional SYM theory dimensionally reduced to one dimension
which describes the low energy dynamics of D0-branes so far, in the
infinite momentum frame,
gives a constructive definition of the
M-theory. This one-dimensional SYM quantum mechanical model is
called M(atrix) theory.
Moreover, by considering toroidal compactification on a circle and
on a two-torus in the manner of Taylor \cite{Taylor},
it leads to a proposal for
a nonperturbative definition of
type IIA superstring theory \cite{DVV,Banks-Seiberg} and
that of type IIB superstring theory
\cite{Banks-Seiberg,Sethi-Susskind}, respectively.
In this paper, we call them IIA and IIB matrix string theories.
The IIA matrix string theory is given by
ten-dimensional SYM theory dimensionally
reduced to two-dimensions, and the string coupling constant corresponds
to the inverse of the SYM coupling. Also, the IIB matrix string theory
is a dimensional reduction of the ten-dimensional SYM theory
to three dimensions. The string coupling is given by a ratio of
lengths of two spatial dimensions of the three-dimensional SYM theory.
In addition, with respect to another nonperturbative definition
of type IIB superstrings \cite{IKKT}, which is in form of
the ten-dimensional SYM theory reduced to a point (zero dimension),
proposed by Ishibashi, Kawai, Kitazawa and Tsuchiya,
it is referred to IKKT model.
Those matrix string theories successfully reproduce some of
known results obtained from an analysis of BPS saturated states
consisting of fundamental strings, D-branes and their bound states
\cite{Banks-Seiberg-Shenker}.
This fact is considered as one of evidences
that the matrix string theories truely are constructive
definitions of string theories. However, with respect to
nonperturbative dynamics of the string theories, in order to
investigate it we cannot help treating nonperturbative dynamics of
the SYM theory sides.
Though understanding of this area is now
in development \cite{Seiberg,PSS},
unfortunately at present it seems to be not powerful enough for
handling the problem.
In this paper, we consider the partition function of the matrix
string theories --- it is one of the most fundamental quantities
reflecting dynamical property of vacua of the theories.
The partition function of the IIA matrix string theory was computed
in strongly coupled limit of the SYM theory
by Kostov and Vanhove \cite{Kostov-Vanhove}.
For the partition function of the IKKT model, the exact result was
obtained by interpreting the theory
as a cohomological field theory
by Moore, Nekrasov and Shatashvili \cite{Moore-Nekrasov-Shatashvili}.
Here, we exactly calculate the partition function of
both of IIA and IIB
matrix string theories by mapping the theories to
cohomological field theories.
In the IIA case our result coincides with the result obtained in the
strongly coupled limit, which thus gives a proof of
{\it exact quasi classics}
discussed in \cite{Kostov-Vanhove}.
Also in the IIB case, our result agrees with the exact calculation for
the IKKT model, which seems to suggest the
equivalence between those two different nonperturbative formulations
of IIB string theory.
The paper is organized as follows. In the next section, we introduce
the IIA and IIB matrix string theories via toroidal compactification
of the M(atrix) theory. We briefly explain how string interactions
emerge from the SYM theory
and show how spinors with correct chirality appear.
In section 3, as a preparation for computing the partition function of
the matrix strings, we evaluate the partition function of
four-dimensional ${\cal N}=4$ $SU(N)$ SYM theory defined on a four-torus
by mapping the theory to a cohomological field theory via
a twisting procedure.
Then in section 4, we exactly compute the partition function of the IIA
matrix string theory, and see that our result coincides with the result
in the infra-red limit by Kostov and Vanhove.
In section 5, for the partition function of the IIB matrix string theory
we perform the calculation in the ten-dimensional IIB limit, and find
the identical result with the partition function of the IKKT model
by Moore, Nekrasov and Shatashvili.
Finally, section 6 is devoted to conclusions.
In Appendix A, we clarify the connection between two-dimensional
${\cal N}=2$ SYM theory and a cohomological field theory used in section 4.
\section{Matrix Strings}
In this section we review some basic properties of the IIA and IIB
matrix string theories, which are derived from the M(atrix) theory by
considering toroidal compactification on $S^1$ and on $T^2$
respectively. In addition, we give an argument
that two spinors of the same space-time chirality emerge
in the IIB matrix string theory, which has not been found
in the literatures.
We start with the M(atrix) theory \cite{BFSS}, whose action
has the same form as low energy effective action of $N$ D0-branes,
that is one-dimensional $U(N)$ SYM theory with 16 supercharges.
The BFSS conjecture is that this action exactly describes M-theory in
the decompactified limit $R_{11} \rightarrow \infty$ by going to
the infinite momentum frame. The infinite momentum frame
means infinite amount of boosting along the 11th direction,
i.e. momentum of the 11th direction becomes
\begin{equation}
p_{11}=\frac{N}{R_{11}}\rightarrow \infty.
\end{equation}
At the same time we must take the limit
$N,\; R_{11} \rightarrow \infty,$ in order that the resulting theory
represents the strongly
coupled limit of type IIA superstring theory.
\subsection{IIA Matrix String Theory}
Let us consider the compactification of one of the transverse
directions, say the 9th, to a circle of the circumference $L,$
denoted by $S^1(L).$
Since in the M(atrix) theory we consider the limit that
the 11th direction is decompactified,
we have ten-dimensional
type IIA string theory with the string length
\begin{equation}
l_s^2=\frac{2\pi l_{11}^3}{L}.
\end{equation}
According to the prescription of Taylor \cite{Taylor},
this IIA string theory is described by two-dimensional ${\cal N} =8 \;
U(N)$ SYM theory, which is a dimensional reduction of
ten-dimensional ${\cal N}=1$ SYM theory \cite{DVV},
\begin{eqnarray}
S_{{\rm IIA-MS}} & = & \frac{1}{g^2}\int dt \int_0^Rd\sigma\;{\rm tr}
\left[-\frac 14F_{\mu\nu}F^{\mu\nu}-\frac 12(D_{\mu}X^I)^2
+i\theta^T(D_t+\Gamma^9D_{\sigma})\theta \right.\nonumber \\
& & \left. +\frac 14 [X^I, X^J]^2
+\theta^T\Gamma^I[X_I,\theta]\right].
\label{actionIIA-MS}
\end{eqnarray}
Here, $F_{\mu\nu}=\partial_{\mu}A_{\nu}-\partial_{\nu}A_{\mu}
-i[A_{\mu},A_{\nu}]$ is a field strength made from two-dimensional
gauge field $A_{\mu},$
$X^I$ ($I=1,\cdots,8$) are Higgs fields and $\theta^{\alpha}$
($\alpha=1,\cdots,16$) are fermions.
We use the convention that
the $\Gamma$-matrices are $16\times 16$ real symmetric matrices
satisfying
\begin{equation}
\{\Gamma^I,\Gamma^J\}=2\delta^{IJ}, \hspace{1cm}
\Gamma^9=\Gamma^1\cdots\Gamma^8.
\end{equation}
$\theta^{\alpha}$'s are decomposed into the spinor and conjugate
spinor representations (${\bf 8_s}\oplus {\bf 8_c}$)
of the rotational group in the transverse directions $SO(8),$
which are characterized by the eigenvalues of $\Gamma^9,$
so called chirality:
\begin{equation}
\theta^{\alpha}=\theta^{\alpha}_++\theta^{\alpha}_-, \hspace{1cm}
\Gamma^9\theta^{\alpha}_{\pm}=\pm\theta^{\alpha}_{\pm},
\label{chiraldecomposition}
\end{equation}
and $X^I$'s transform as the vector ${\bf 8_v}.$
The spatial coordinate $\sigma,$ which arises as performing the
compactification, takes a value on the circle dual to
$S^1(L)$: $0\leq\sigma\leq R,$ where\footnote{
The relation between parameters in the matrix string theory and those
in SYM theory, eqs. (\ref{RLrelation}) and (\ref{IIAggs}) in the IIA case
as well as eqs. (\ref{*}), (\ref{**}) and (\ref{IIBcoupling})
in the IIB case, can be derived either by tracing the procedure by
Taylor \cite{Taylor} or
by employing another argument in ref. \cite{FHRS}.}
\begin{equation}
R=(2\pi)^2\frac{l_s^2}{L}.
\label{RLrelation}
\end{equation}
The coupling constant of the SYM theory $g$ is related to the string
coupling $g_s$ as
\begin{equation}
g^2=\frac{2\pi}{(Rg_s)^2}.
\label{IIAggs}
\end{equation}
Weakly coupled strings are recovered by considering the limit
$g_s\rightarrow 0$ or equivalently the infra-red limit of the
SYM theory \cite{DVV}. In this situation, the theory is
described by the eigenvalues of the simultaneously
diagonalizable configurations of $X^I$ and $\theta^{\alpha}$:
\begin{equation}
\tilde{X^I} = {\rm diag}(x^I_1,\cdots,x^I_N), \hspace{1cm}
\tilde{\theta^{\alpha}} =
{\rm diag}(\theta^{\alpha}_1,\cdots, \theta^{\alpha}_N)
\end{equation}
where
\begin{equation}
X^I=V\tilde{X^I}V^{\dagger}, \hspace{1cm}
\theta^{\alpha}=V\tilde{\theta^{\alpha}}V^{\dagger}.
\label{diagonalpart}
\end{equation}
The angular variables $V\in U(N)$ and nontrivial
configurations of the gauge field yielding
non-zero curvature $F_{\mu\nu}$ are
energetically decoupled from the theory in the infra-red limit,
and the action (\ref{actionIIA-MS}) reduces to the action of
multiple Green-Schwarz superstrings in the light-cone gauge.
Then the gauge field can take the pure gauge configuration
\begin{equation}
A_{\mu}=iV\partial_{\mu}V^{\dagger}.
\end{equation}
Also, the $x^I_i$ and $\theta^{\alpha}_i$ can represent
strings of various lengths by considering the multi-valued
configuration:
\begin{eqnarray}
& & \tilde{X^I}(t,\sigma+R) =
g^{\dagger}\tilde{X^I}(t,\sigma)g, \hspace{1cm}
\tilde{\theta^{\alpha}}(t,\sigma+R) =
g^{\dagger}\tilde{\theta^{\alpha}}(t,\sigma)g, \nonumber \\
& & V(t,\sigma+R) = V(t,\sigma)g,
\label{BCofdiagonalpart}
\end{eqnarray}
where $g$ is an element of the Weyl group of $U(N),$ i.e. the
permutation group $S_N,$ which permutes the $N$ eigenvalues.
In going around the $\sigma$-direction, the eigenvalues are
interchanged by the action of $g$ in eqs. (\ref{BCofdiagonalpart}),
and as a result they form cycles of various
lengths corresponding to permutation cycles in $g.$
Each cycle is interpreted as a single closed string, and thus
for $g\in S_N$
consisting of $N_n$ $n$-cycles
(satisfying $N=\sum_nnN_n$) one has $N_n$ strings with length
$n.$
Note that the total matrices
$X^I,$ $\theta^{\alpha}$
in eqs. (\ref{diagonalpart}) and the gauge field $A_{\mu},$
which appear in the SYM theory,
remain single-valued although we consider
the case that the variables of string coordinates
$\tilde{X^I}$ and $\tilde{\theta^{\alpha}}$ are multi-valued
as in eqs. (\ref{BCofdiagonalpart}),
which is argued in ref. \cite{Wynter}.
For the simplest
$N=2$ case, the multi-valued configuration is represented by
\begin{equation}
\tilde{X^I}=\left(\begin{array}{cc} x^I_1 & 0 \\
0 & x^I_2
\end{array}\right), \hspace{1cm}
\tilde{\theta^{\alpha}}=\left(\begin{array}{cc}
\theta^{\alpha}_1 & 0 \\
0 & \theta^{\alpha}_2
\end{array}\right), \hspace{1cm}
V=\frac{1}{\sqrt{2}}\left(\begin{array}{cc}
e^{i\pi\sigma/R} & -e^{i\pi\sigma/R} \\
1 & 1
\end{array}\right),
\end{equation}
where
$$
x^I_1(t,\sigma+R)=x^I_2(t,\sigma), \hspace{1cm}
x^I_2(t,\sigma+R)=x^I_1(t,\sigma),
$$
and $\theta^{\alpha}_i$ satisfies the same boundary condition.
The matrices $X^I,$ $\theta^{\alpha}$ and $A_{\mu}$ are
single-valued. In particular, the gauge field is
\begin{equation}
A_t=0, \hspace{1cm}
A_{\sigma}=\frac 12 \left(\begin{array}{cc} 1 & 0 \\
0 & 0
\end{array}\right),
\end{equation}
which cannot be eliminated by any single-valued gauge
transformation.
As discussed in \cite {DVV},
relaxing the strict limit $g_s=0,$ the number of the
strings are no longer conserved. We can interpret
this as a result of string interactions
(splitting or joining of one or two strings) which occur
when the two sets of eigenvalues $\{x^I_i\}_{I=1,\cdots,8}$ and
$\{x^I_j\}_{I=1,\cdots,8}$ coincide.
Whenever the interaction occurs, the boundary condition of
$\tilde{X^I},$ $\tilde{\theta^{\alpha}}$ and $V$ is changed,
i.e. a branch point appears on a cylinder coordinated by
$t$ and $\sigma.$
We should remark that as discussed in \cite{GHV},
in spite of this singular configuration
of the diagonal variables, for a neighborhood of the
branch point the corresponding total matrices
exist as a smooth and single-valued
configuration of the SYM theory.
In the section 4
of ref. \cite{GHV} such a configuration is constructed
in the $N=2$ case, which is
given by a solution of the dimensionally reduced version
of the self-dual equation of four-dimensional Yang-Mills
theory to two dimensions:
\begin{eqnarray}
& & [X_1,X_2]=\frac{i}{g^2}F_{t\sigma}, \label{selfdualeq} \\
& & D_tX_1=-D_{\sigma}X_2, \nonumber \\
& & D_tX_2=D_{\sigma}X_1,
\end{eqnarray}
and the all other fields are set to zero.
Compared to the configuration in the strict limit ($g^2=\infty$)
that is $[X_1,X_2]=0,$ the $O(1/g^2)$-correction in
eq. (\ref{selfdualeq}) works well to make the configuration
in the SYM theory smooth\footnote{An extension of this argument
in the case of general $N$ is discussed in refs. \cite{BBN}.
I thank L. Bonora for informing me of those literatures.}.
\subsection{IIB Matrix String Theory}
When compactifying IIA string theory to a circle and taking
T-dual to the circle, we have IIB string theory on $S^1.$
In the limit of shrinking the circle of the IIA theory, the $S^1$ of
the IIB theory is decompactified, thus we have IIB string theory in
ten dimensions. From the view point of M-theory, IIA theory
on $S^1$ means M-theory on $T^2.$ We denote the size of the
two-torus by $L_1$ and $L_2,$ i.e. $T^2=S^1(L_1)\times S^1(L_2).$
Since there is no distinctive meaning between the two $S^1$'s,
now we have
two ways to obtain IIB theory
corresponding to shrinking either $L_1$ or
$L_2.$ From the analysis based on low energy effective
theories, two IIB theories we obtain as the result are believed to be
equivalent and connected by S-duality \cite{Aspinwall,Schwarz}.
In this way, the M-theory perspective yields a geometrical
interpretation to the S-duality in type IIB string theory.
Let us consider this operation in the M(atrix) theory.
Then we obtain three-dimensional ${\cal N}=8$ $U(N)$ SYM theory,
which is dimensionally reduced from the ten-dimensional theory, as
a matrix model corresponding to IIB theory on $S^1$
\cite{Banks-Seiberg,Sethi-Susskind}:
\begin{eqnarray}
S_{{\rm IIB-MS}} & = & \frac{1}{g^2}\int dt \int_0^{R_1}d\sigma_1
\int_0^{R_2}d\sigma_2\;{\rm tr}
\left[-\frac 14F_{\mu\nu}F^{\mu\nu}-\frac 12(D_{\mu}X^I)^2
\right.\nonumber \\
& & \left.+i
\theta^T(D_t+\Gamma^8D_{\sigma_1}+\Gamma^9D_{\sigma_2})
\theta
+\frac 14 [X^I, X^J]^2
+\theta^T\Gamma^I[X_I,\theta]\right].
\label{actionIIB-MS}
\end{eqnarray}
Here we compactified the 8th and 9th transverse directions to
the above mentioned (rectangular) two-torus. The size of the
spatial directions ($\sigma_1$ and $\sigma_2$)
which the SYM theory is defined on is related
by T-duality to the two-torus,
\begin{equation}
R_i=(2\pi)^2\frac{l_s^2}{L_i} \hspace{1cm} (i=1,2),
\label{*}
\end{equation}
and the SYM coupling are given by
\begin{equation}
g^2=\frac{R_1R_2R_{11}}{(2\pi)^2l_s^4}.
\label{**}
\end{equation}
It is remarkable that the string coupling is given by
a ratio of the two
lengths of the torus:
\begin{equation}
g_s=\frac{R_1}{R_2}=\frac{L_2}{L_1},
\label{IIBcoupling}
\end{equation}
which gives the geometrical understanding of S-duality
also in the matrix string level.
In the vanishing torus limit $L_1, L_2\rightarrow 0,$ a new dimension
opens up and becomes
decompactified, so this limit with the ratio $L_2/L_1$
fixed is considered to give type IIB theory of the coupling $g_s$
determined by eq. (\ref{IIBcoupling}) \cite{Sethi-Susskind,FHRS}.
Note that the abelian part of
the field strength $F_{\mu\nu}$ made from
three-dimensional gauge field $A_{\mu}$ reduces to a single
scalar field via duality transformation.
Although the manifest symmetry of the action (\ref{actionIIB-MS}) is
$SO(7),$ it is considered that
this scalar and the seven Higgs fields $X^I$
($I=1,\cdots,7$)
together belong to ${\bf 8_v}$ in $SO(8)$
in the ten-dimensional IIB limit
\begin{equation}
R_1R_2\rightarrow \infty, \hspace{1cm} g_s :{\rm fixed}
\label{10DIIBlimit}
\end{equation}
by the argument for BPS states \cite{FHRS} and by the analysis
of the moduli space of the three-dimensional SYM theory
\cite{Banks-Seiberg,Seiberg}.
The fermions $\theta^{\alpha}$ ($\alpha=1,\cdots,16$) are
to represent
two space-time spinors of the same chirality in
the IIB theory. In fact, after decomposing by the eigenvalues of
$\Gamma^9$ as in eq. (\ref{chiraldecomposition}), we put
\begin{equation}
\psi^{\alpha}_+=\Gamma^8\theta^{\alpha}_-.
\label{psi+}
\end{equation}
Then fermion part of the lagrangian density takes the form
\begin{eqnarray}
& & {\rm tr} \;[i\theta^T_+(D_t+D_{\sigma_2})\theta_+
+i\psi^T_+(D_t-D_{\sigma_2})\psi_+
+i\theta^T_+D_{\sigma_1}\psi_+
+i\psi^T_+D_{\sigma_1}\theta_+ \nonumber \\
& & \; +\theta^T_+\Gamma^I\Gamma^8[X^I,\psi_+]
-\psi^T_+\Gamma^I\Gamma^8[X^I,\theta_+]].
\label{IIBfermionpart}
\end{eqnarray}
Both of $\theta_+$ and $\psi_+$ are spinors with respect to
the manifest symmetry $SO(7)$
of the same eigenvalue (+1) of $\Gamma^9.$
In the weakly coupled limit $g_s\ll 1,$ which means $R_1\ll R_2,$
nonzero modes of $\partial_{\sigma_1}$
energetically decouple, and $\theta_+$ and $\psi_+$ represent
the two ${\bf 8_s}$ spinors in IIB theory.
Also, in the strongly coupled limit $R_2\ll R_1$,
it turns out that
$\xi_+=(\theta_++\psi_+)/\sqrt 2$ and
$\eta_+=(\theta_+-\psi_+)/\sqrt 2$ become spinors desired in
IIB theory\footnote{
In IIA matrix string theory,
by the same replacement (\ref{psi+}), we seem to have
chiral IIB theory, but it is not correct.
The IIA matrix string theory has the manifest symmetry $SO(8).$
Note that $\psi_+$ is not a spinor under the $SO(8)$ due to
the $\Gamma^8$ factor in eq. (\ref{psi+}).}.
In this way, we can see that
in both of the two limits related via S-duality chiral spinors are
correctly
reproduced in the IIB matrix string theory.
\section{Partition Function of four-dimensional ${\cal N}=4$ $SU(N)$
Super Yang-Mills Theory}
Here we calculate the partition function of
${\cal N}=4$ $SU(N)$ SYM theory on four-torus $T^4$
by mapping the theory to a cohomological field theory.
The dimensional reduced version of this argument
is used in later computations
of the partition function of IIA and IIB matrix string theories.
${\cal N}=4$ $SU(N)$ SYM theory on a flat four-dimensional space-time
is given by the following lagrangian density
defined on the ${\cal N}=1$
superspace $(x,\theta,\bar{\theta})$:
\begin{eqnarray}
{\cal L}_{{\cal N}=4} & = & \frac{1}{16g^2}{\rm tr}
\left(W^{\alpha}W_{\alpha}|_{\theta\theta}+
\bar{W}_{\dot{\alpha}}
\bar{W}^{\dot{\alpha}}|_{\bar{\theta}\bar{\theta}}\right)
\nonumber \\
& & +\frac{1}{g^2}{\rm tr} \left(\Phi_1^{\dagger}e^V\Phi_1+
\Phi_2^{\dagger}e^V\Phi_2+
\Phi_3^{\dagger}e^V\Phi_3\right)
|_{\theta\theta\bar{\theta}\bar{\theta}} \nonumber \\
& & +\frac{1}{\sqrt{2}g^2}{\rm tr} \left(
\Phi_1[\Phi_2,\Phi_3]|_{\theta\theta}+
\Phi_1^{\dagger}[\Phi_3^{\dagger},\Phi_2^{\dagger}]
|_{\bar{\theta}\bar{\theta}}\right).
\label{4DN=4}
\end{eqnarray}
In this section, we use the notation in ref. \cite{Wess-Bagger}.
The vector superfield $V$ represents a multiplet containing
a gauge field $A_m$ and
a complex two-component gauge fermion $\lambda,$
and the chiral superfield
$\Phi_s$ ($s=1,2,3$) contains a multiplet of a complex Higgs scalar
$B_s$ and a Higgsino $\psi_s.$
All the fields belong to the adjoint representation of $SU(N),$
``${\rm tr}$'' denotes the trace in the fundamental representation.
The theory has the internal symmetry group $SU(4)_I,$
under which the fermions ($\lambda$ and $\psi_s$'s) transform
together as {\bf 4} and $B_s$'s as {\bf 6}.
Also, the supercharges $Q_{\alpha}^v,$ $\bar{Q}_{v \dot{\alpha}}$
as {\bf 4}, ${\bf \bar{4}}$ respectively, where
$v\;(=1,\cdots,4)$ is the $SU(4)_I$ index and
$\alpha,$ $\dot{\alpha}$ are Lorentz indices belonging to
$SU(2)_L,$ $SU(2)_R.$
In terms of the component fields, the lagrangian takes the form
\begin{eqnarray}
{\cal L}_{{\cal N}=4} & = & \frac{1}{g^2}{\rm tr} \left[
-\frac 14 F^{mn}F_{mn}-i\bar{\lambda}\bar{\sigma}^mD_m\lambda
\frac 12 D^2\right] \nonumber \\
& & \left.+\frac{1}{g^2}\sum_{s=1}^3{\rm tr}\right[
F_s^{\dagger}F_s-(D^mB_s)^{\dagger}(D_mB_s)
-i\bar{\psi}_s\bar{\sigma}^mD_m\psi_s \nonumber \\
& & \hspace{15mm}
\left. -\frac{i}{\sqrt 2}B_s^{\dagger}[\lambda,\psi_s]
+\frac{i}{\sqrt 2}B_s[\bar{\lambda},\bar{\psi}_s]
+\frac 12 D[B_s,B_s^{\dagger}]\right] \nonumber \\
& & +\frac{1}{\sqrt 2 g^2}{\rm tr} (F_1[B_2,B_3]+F_2[B_3,B_1]+F_3[B_1,B_2]
\nonumber \\
& & \hspace{15mm}
-B_1[\psi_2,\psi_3]-B_2[\psi_3,\psi_1]-B_3[\psi_1,\psi_2]
\nonumber \\
& & \hspace{15mm} +{\rm h. c.}),
\end{eqnarray}
where
$F_s$ and $D$ are auxiliary fields appearing in $\Phi_s$ and $V.$
\subsection{Twisting of ${\cal N}=4$ Super Yang-Mills Theory}
The twisting procedure, which was first introduced by Witten
\cite{Witten3}, gives a systematic tool for
constructing a cohomological
field theory from the original physical theory.
Here we briefly explain a twisting procedure of $N=4$ SYM theory
adopted by
Vafa and Witten \cite{Vafa-Witten}.
First, we consider a general four-manifold on which the ${\cal N}=4$
SYM theory is defined. The holonomy group of this manifold is
$SO(4)=SU(2)_L\times SU(2)_R,$ that is a gauged symmetry with the
spin connection being the (external) gauge field.
Adding the global $SU(4)_I,$ we concentrate
the symmetry group of the theory
$$
H=SU(2)_L\times SU(2)_R\times SU(4)_I.
$$
The twisting by Vafa and Witten is performed as follows.
Consider the subgroup of the internal symmetry group $SU(4)_I$:
$$
SU(4)_I\supset SO(4)=SU(2)_F\times SU(2)_{F'},
$$
then replace the action of $SU(2)_L$ by the diagonal sum
$SU(2)'_L=SU(2)_L\oplus SU(2)_{F'}.$
Thus the twisted symmetry group becomes
$$
H'=SU(2)'_L\times SU(2)_R\times SU(2)_F.
$$
Under $H',$ the supercharges split up as
\begin{eqnarray*}
& & Q_{\alpha}^v=Q_{\alpha}^{ij}\rightarrow
Q_{\alpha}^{i\beta}=\left\{\begin{array}{l}
Q_{\alpha}^{i\alpha}\equiv Q^i \\
\frac 12 \varepsilon_{\gamma (\beta}Q_{\alpha)}^{i\gamma}
\equiv Q_{\alpha\beta}^i
\end{array}\right. \\
& & \bar{Q}_{v \dot{\alpha}}=\bar{Q}_{ij\dot{\alpha}}\rightarrow
\bar{Q}_{i\beta\dot{\alpha}},
\end{eqnarray*}
where $i$ and $j$ are $SU(2)_F$ and $SU(2)_{F'}$ indices
respectively, also $j$ is converted to $\beta$ by the twisting.
Here we have the two scalar supercharges $Q^i$ ($i=1,2$), which
are nilpotent
\begin{equation}
\{Q^i, Q^j\}=0
\end{equation}
as seen from the supersymmetry algebra of the original
supercharges
\begin{eqnarray}
& & \{Q_{\alpha}^v,\bar{Q}_{w \dot{\alpha}}\}=
2\sigma^m_{\alpha\dot{\alpha}}\delta^v_w P_m, \nonumber \\
& & \{Q_{\alpha}^v,Q_{\beta}^w\}=
\{\bar{Q}_{v\dot{\alpha}},\bar{Q}_{w\dot{\beta}}\}=0.
\label{***}
\end{eqnarray}
We can construct a cohomological field theory based on
each $Q^i$'s. The $Q^i$'s are related to the original supercharges as
\begin{eqnarray}
Q^1 & = & Q_1^{11}+Q_2^{12}=Q_{\alpha=1}^{v=1}+Q_{\alpha=2}^{v=2},
\nonumber \\
Q^2 & = & Q_1^{21}+Q_2^{22}=Q_{\alpha=1}^{v=3}+Q_{\alpha=2}^{v=4}.
\label{Qrelation}
\end{eqnarray}
Next, in the case of four-dimensional K\"{a}hler manifolds,
the number of the nilpotent scalar supercharges become doubled
as we will see\footnote{This property was explored first by Witten
\cite{Witten4} in the case of ${\cal N}=2$ SYM theory.}.
A four-dimensional K\"{a}hler manifold has the reduced
holonomy group $U(1)_L\times SU(2)_R,$ where $U(1)_L$ is a certain
subgroup of $SU(2)_L.$ In this case, twisting is done by replacing
the $U(1)_L$ by the diagonal sum
$U(1)'_L=U(1)_L\oplus U(1)_{F'}.$
Here, $U(1)_{F'}$ is a subgroup of $SU(2)_{F'}.$
We assign the $U(1)_L$ charge $+1$ to $Q_{\alpha=1}^v$ and $-1$
to $Q_{\alpha=2}^v,$ and give a similar assignment of the
$U(1)_{F'}$ charge ($+1$ to the $j=1$ index and $-1$ to the $j=2$).
Then by twisting, the four $Q_{\alpha}^v$'s
in eqs. (\ref{Qrelation}) are to have zero $U(1)'_L$ charge, that is,
they all become scalar supercharges, since
$Q_1^{i1}=Q_{12}^i$ and $Q_2^{i2}=-Q_{21}^i.$
It is clear that they are nilpotent.
Note that one of the four charges is a (chiral) generator of
the manifest ${\cal N}=1$ supersymmetry in eq. (\ref{4DN=4}).
We will use that charge, denoted by $Q,$ to construct
a cohomological field theory in the next subsection.
Since the twisting changes the coupling of the fields with
the spin connection \cite{Lab-Marino1},
for a general four-manifold the twisted theory
differs from the original theory. However, on a manifold
with the flat metric ($T^4$) or more generally
on a hyper-K\"{a}hler manifold for which the twisting is a trivial
operation
owing to its holonomy group $SU(2)_R,$ the twisted theory coincides
with the original theory. Now we will investigate the $T^4$ case.
\subsection{Partition function of ${\cal N}=4$ SYM on $T^4$}
The ${\cal N}=4$ $SU(N)$ SYM theory defined on $T^4$
can be seen as a cohomological field theory by (trivial) twisting,
because of the reason mentioned above.
In the action $S_{{\cal N}=4}$ (\ref{4DN=4}), since $Q$ acts as
$\frac{\partial}{\partial\theta^{\alpha}}=\int d\theta^{\alpha}$
(up to total derivative terms) for suitable $\alpha,$
the term with $|_{\theta\theta}$ are written in the $Q$-exact form
$\{Q,\cdots\}.$ Also, the term ${\rm tr}\bar{W}_{\dot{\alpha}}
\bar{W}^{\dot{\alpha}}|_{\bar{\theta}\bar{\theta}}$
is equal to ${\rm tr} W^{\alpha}W_{\alpha}|_{\theta\theta}$ modulo
the total derivative term ${\rm tr} F\wedge F,$
so it is of the $Q$-exact form.
Here we consider the periodic boundary condition
case on $T^4,$ or equivalently the sector of zero 't Hooft
discrete magnetic flux, those total derivative terms can be
discarded.
The term
$
{\rm tr} \;\Phi_1^{\dagger}[\Phi_3^{\dagger},\Phi_2^{\dagger}]
|_{\bar{\theta}\bar{\theta}}
$
can not be written as the $Q$-exact form,
but it is $Q$-invariant because $Q$ is a
part of the manifest ${\cal N}=1$ supersymmetric transformation.
Thus the action
can be written as
\begin{equation}
S_{{\cal N}=4}=\{Q,\cdots\}+{\cal O}^{(0)},
\end{equation}
where
\begin{equation}
{\cal O}^{(0)}= \int d^4x \frac{1}{\sqrt 2g^2}{\rm tr} \;
\Phi_1^{\dagger}[\Phi_3^{\dagger},\Phi_2^{\dagger}]
|_{\bar{\theta}\bar{\theta}}
\end{equation}
is a $Q$-invariant operator and the superscript 0 stands for
the $U(1)$-charge defined by the rotation
\begin{eqnarray}
\Phi_s(\theta,\bar{\theta},x) & \rightarrow & e^{\frac 23i\beta}
\Phi_s(e^{-i\beta}\theta,e^{i\beta}\bar{\theta},x) \nonumber \\
W^{\alpha}(\theta,\bar{\theta},x) & \rightarrow &
e^{i\beta}W^{\alpha}(e^{-i\beta}\theta,e^{i\beta}\bar{\theta},x)
\label{U(1)charge}
\end{eqnarray}
with $\beta$ being a real parameter.
The charge of the component fields can be read off from eqs.
(\ref{U(1)charge}) as
$-\frac 23,\;+\frac 13,\;+\frac 43,\;0,\;-1,\;0$ for
$B_s,\;\psi_s,\;F_s,\;A_m,\;\lambda,\;D,$ respectively.
Note that the action has the symmetry under this rotation.
Now the problem can be made more tractable by considering the following
mass perturbation
\begin{equation}
\Delta S = \int d^4x \frac{1}{\sqrt 2 g^2} \left(-\frac m2\right) {\rm tr} \;
[(\Phi_1^2+\Phi_2^2+\Phi_3^2)|_{\theta\theta}
+(\Phi_1^{\dagger 2}+\Phi_2^{\dagger 2}+\Phi_3^{\dagger 2})
|_{\bar{\theta}\bar{\theta}}].
\label{masspert}
\end{equation}
Using the same argument as above,
the first term is written in the form $\{Q,\cdots\},$ and the
second term is a $Q$-invariant operator with the $U(1)$-charge
$-\frac 23$ denoted by ${\cal O}^{(-2/3)}.$
In cohomological field theories, the perturbation
by the $Q$-exact operators does not alter the theory,
and the partition function is invariant under the perturbation of
${\cal O}^{(-2/3)}$ if the $U(1)$-symmetry remains left even
in the quantum level. In fact, it is so since there is no anomaly
in ${\cal N}=4$ theory. Thus we can obtain the answer
by computing the partition function of
the mass perturbed system.
Here, it is remarked that the
$U(1)$-symmetry is different from the ghost number symmetry usually
used in cohomological field theories. In fact, it is seen that
in the cohomological field theory
made from ${\cal N}=4$ SYM theory the ghost number symmetry corresponds to
the following rotation \cite{Vafa-Witten,Lab-Marino2}
\begin{eqnarray*}
\Phi_1(\theta,\bar{\theta},x) & \rightarrow &
\Phi_1(e^{-i\beta}\theta,e^{i\beta},x) \\
\Phi_2(\theta,\bar{\theta},x) & \rightarrow &
\Phi_2(e^{-i\beta}\theta,e^{i\beta},x) \\
\Phi_3(\theta,\bar{\theta},x) & \rightarrow &
e^{2i\beta}\Phi_3(e^{-i\beta}\theta,e^{i\beta},x) \\
W^{\alpha}(\theta,\bar{\theta},x) & \rightarrow &
e^{i\beta}W^{\alpha}(e^{-i\beta}\theta,e^{i\beta}\bar{\theta},x).
\end{eqnarray*}
Vafa and Witten \cite{Vafa-Witten} discussed the invariance of the
partition function under
the mass perturbation based on this ghost number symmetry.
Our argument presents another proof of the invariance.
We should remark that cohomological field theories have an important
feature that the contribution of the path integration localizes
in configurations of the classical vacua. In the mass perturbed system,
the classical vacua are given by the solutions of these equations:
\begin{eqnarray}
& & [ B_1, B_2]=m B_3, \nonumber \\
& & [ B_2, B_3]=m B_1, \nonumber \\
& & [ B_3, B_1]=m B_2
\label{classicalvacua1}
\end{eqnarray}
with
\begin{equation}
\sum_{s=1}^3[B_s,B_s^{\dagger}]=0.
\label{classicalvacua2}
\end{equation}
A special property of the perturbation is that three equations in
(\ref{classicalvacua1}) have the form of $SU(2)$ algebra.
So, the solution $\frac im B_s$ is given by the generator of
$N$-dimensional representation
of $SU(2).$ At the same time, this solution satisfies eq.
(\ref{classicalvacua2}) also, because $B_s$'s are anti-hermitian.
Since we must consider the various $N$-dimensional representations,
both of reducible and irreducible, in general the solution takes the
form
\begin{equation}
\frac im B_s = \left[ \begin{array}{cccc}
L_s^{(a_1)} & & & \\
& L_s^{(a_2)} & & \\
& & \ddots & \\
& & & L_s^{(a_l)}
\end{array} \right],
\label{generalsol}
\end{equation}
where $L_s^{(a)}$ is a generator of $a$-dimensional
irreducible representation of $SU(2),$ and $a_1+a_2+\cdots+a_l=N.$
We remark that all the solutions of the form (\ref{generalsol}) do not
contribute to the partition function, but only the $a_1=a_2=\cdots=a_l$
case does. In fact, considering the $l=2$ case, it is easy to see
that the unbroken gauge group of the solution (\ref{generalsol})
contains a following $U(1)$-generator
\begin{equation}
\left[ \begin{array}{cc}
e_1{\bf 1}_{a_1} & 0 \\
0 & e_2{\bf 1}_{a_2}
\end{array} \right],
\end{equation}
where $e_1$ and $e_2$ are real parameters satisfying $e_1a_1+e_2a_2=0.$
Let us consider the $a_1\neq a_2$ case.
The interactions are all in the form of commutators,
so this $U(1)$ mode
is free and massless, which implies that there exists
a fermion zero-mode corresponding to the $U(1).$ Thus its contribution
to the partition function vanishes. However, when $a_1=a_2,$
the $B_s$ is written as
\begin{equation}
\frac im B_s={\bf 1}_2\otimes L_s^{(a)},
\end{equation}
where we put $a_1=a_2\equiv a.$
Then it is noted that
the unbroken gauge group is enhanced to $SU(2)\otimes {\bf Z}_a.$
This mode is not free and thus can give a certain contribution to
the partition function, which amounts to a
${\cal N}=1$ $SU(2)\otimes {\bf Z}_a$
SYM theory. For the case of generic $l,$ a similar argument goes on
and it can be seen that the vacuum which can contribute to the
partition function is a ${\cal N}=1$ $SU(l)\otimes {\bf Z}_a$ SYM theory
with $al=N.$ In the consequence the partition function of
the ${\cal N}=4$ theory is represented as a sum of the partition function
of the various ${\cal N}=1$ theories:
\begin{equation}
Z_{SU(N)}^{D=4,\; {\cal N}=4}(T^4)=
\sum_{al=N}Z_{SU(l)\otimes {\bf Z}_a}^{D=4,\;{\cal N}=1}(T^4).
\label{N=4partitionfunction1}
\end{equation}
Further, since the ${\bf Z}_a$ factor in the gauge group
$SU(l)\otimes {\bf Z}_a$ implies a summation of the flat ${\bf Z}_a$
bundle, it yields a factor $a^{4-1}=a^3,$
where the power 4 comes from summing up the flat bundle for
each $A_m$'s and the ($-1$) from dividing by the gauge group ${\bf Z}_a.$
Thus we have
\begin{equation}
Z_{SU(N)}^{D=4,\; {\cal N}=4}(T^4)=
\sum_{al=N}a^3Z_{SU(l)}^{D=4,\; {\cal N}=1}(T^4).
\label{N=4partitionfunction2}
\end{equation}
Here we should note that the partition function of the ${\cal N}=1$ theory
in the right hand side
is given by its Witten index. In the Hamiltonian formalism,
the partition function is written as
\begin{equation}
{\rm Tr}(-1)^Fe^{-\beta H},
\label{trform}
\end{equation}
where the $(-1)^F$-factor is included
in order to impose the periodic boundary condition on the fermion fields.
Recall that there is
no Higgs field in the ${\cal N}=1$ theory,
which implies that there exists no continuous zero-mode
in the theory on $T^4.$
Spectra appearing in the theory are all discrete.
Then in eq. (\ref{trform}), the contribution of
every supersymmetric pair with non-zero
energy is precisely cancelled, and only the zero-energy states
can contribute.
So eq. (\ref{trform}) is independent of $\beta,$ and it coincides
with the Witten index
\begin{equation}
I\equiv\lim_{\beta\rightarrow\infty}{\rm Tr}(-1)^Fe^{-\beta H},
\end{equation}
whose value of the ${\cal N}=1$ $SU(l)$ SYM theory is known to be $l$
\cite{Witten5}. Plugging the above formulas, we obtain the answer
\begin{equation}
Z_{SU(N)}^{D=4,\; {\cal N}=4}(T^4)=\sum_{al=N}a^3l=N\sum_{a|N}a^2,
\end{equation}
where the summation of $a$ in the right hand side is taken over
the divisors of $N.$
Also, for the gauge group $SU(N)/{\bf Z}_N$ we can obtain the
partition function of the sector of zero 't Hooft magnetic flux
applying the above argument to the ${\bf Z}_N$-factor
\begin{equation}
Z_{SU(N)/{\bf Z}_N}^{D=4,\; {\cal N}=4}(T^4)=
\frac{1}{N^3}Z_{SU(N)}^{D=4,\; {\cal N}=4}(T^4)=
\sum_{a|N}\frac{1}{a^2}.
\end{equation}
\section{IIA Matrix String Partition Function}
Here we calculate the partition function of IIA matrix string
theory and compare the result of Kostov and Vanhove \cite{Kostov-Vanhove}
which has been derived in the case of the strong coupling limit
$g\rightarrow\infty.$
We have seen in section 2 that the SYM fields in periodic boundary
condition can reproduce the second quantized superstrings with
interactions. Thus, considering the two-dimensional SYM theory with
periodic boundary condition as the IIA matrix string theory, we evaluate
the partition function of this theory.
Now we have to take care of the $U(1)$ part of the gauge group $U(N)$
before doing the calculation. We consider the gauge group $U(N)$
in the factorized form
\begin{equation}
U(N)=U(1)\times (SU(N)/{\bf Z}_N).
\label{decomposeU(1)}
\end{equation}
The meaning of the $U(1)$ part in the IIA matrix string theory,
whose field contents are $(A_{\mu},X,\theta),$ is as follows.
The $U(1)$ part of $X$ and $\theta$ represents the center of mass
coordinates of the strings in transverse directions. We fix it from
the translational invariance. On the other hand, the $U(1)$ part of
$A_{\mu}$ is related to the number of D-particles, so we have to take
into account this. More precisely, as is discussed
in \cite{DVV}, the $U(1)$ electric flux corresponds to the number of
D-particles:
\begin{equation}
q=\frac{1}{2\pi}\int_0^Rd\sigma E^{U(1)}\in {\bf Z}.
\end{equation}
Corresponding to eq. (\ref{decomposeU(1)}), the gauge field $A_{\mu}$
is decomposed as
\begin{equation}
A_{\mu}=A^{U(1)}_{\mu}T^{U(1)}+A_{\mu}^aT^a,
\end{equation}
where $T^{U(1)}=\frac 1N{\bf 1}_N,$ and $T^a$ ($a=1,\cdots,N^2-1$) is a
generator of $SU(N).$ Since the interaction in the SYM theory
appears in the form
of commutators, the $U(1)$ gauge part decouples and thus the
partition function of the IIA matrix string theory becomes
\begin{equation}
Z_{{\rm IIA-MS}}=\left(\int
\frac{{\cal D}A^{U(1)}_{\mu}}{{\rm Vol}(U(1))}
e^{-S^{U(1)}}\right)Z_{SU(N)/{\bf Z}_N}^{D=2,\;{\cal N}=8}(T^2),
\label{ZIIAMS}
\end{equation}
where the Wick rotation was performed. The two-torus,
where the SYM theory is defined on, is a rectangular one with the size
$T\times R.$ The action of the $U(1)$ gauge part is given by
$$
S^{U(1)}=\frac{1}{Ng^2}\int d^2\sigma\frac 14
F^{U(1)}_{\mu\nu}F^{U(1)}_{\mu\nu}.
$$
Let us evaluate the first factor in eq. (\ref{ZIIAMS}) by taking
the $A^{U(1)}_0=0$ gauge.
Then, the Gauss law constraint means that
$A^{U(1)}_{\sigma}$ is independent of $\sigma.$ Also, considering the
Wilson loop wrapping around the $\sigma$ direction, we see that
$\theta(\tau)\equiv RA^{U(1)}_{\sigma}$ is an angular variable whose
conjugate momentum $p$ is quantized to an integer. Employing the
variable $\theta(\tau)$ and translating to the Hamiltonian form,
we can compute the first factor as
\begin{eqnarray}
\int \frac{{\cal D}A^{U(1)}_{\mu}}{{\rm Vol}(U(1))}\;e^{-S^{U(1)}}
& = & \int_{\theta(T)=\theta(0)}{\cal D} \theta (\tau)
\;e^{-\frac 12 \frac{1}{Ng^2R}\int_0^Td\tau \dot{\theta}(\tau)^2}
\nonumber \\
& = & {\rm Tr} \left( e^{-T\frac 12 Ng^2Rp^2}\right) \nonumber \\
& = & \sum_{p\in {\bf Z}}e^{-\frac{RT}{2}Ng^2p^2}.
\label{firstfactor}
\end{eqnarray}
\subsection{$Z_{SU(N)/{\bf Z}_N}^{D=2,\;{\cal N}=8}(T^2)$}
Now our remaining task is evaluating the second factor of
eq. (\ref{ZIIAMS}), which is performed by considering the dimensionally
reduced version of the analysis in section 3\footnote{
Since in section 3 the supersymmetry algebra (\ref{***}) has no central
charges, here we are to consider the ${\cal N}=8$ supersymmetry algebra
without central charges as the result of the dimensional reduction.
In fact, the central charges are written in the form of total derivatives,
so they do not appear under the periodic boundary condition.
In the infinite two-dimensional space, central charges exist and
represent the topological charges which characterize stable solitonic
modes in the theory. On the contrary, in the two-torus
with the finite size, such solitonic modes do not exist stably.
In this case, it can be considered that with respect to the sector of
the zero total charge, dynamical degrees of freedom of the modes are
contained in the theory, i.e. in the configurations
which the path integration is performed over,
and that they arise as metastable states
when the sizes of the two-torus becomes large enough.
Of course, a similar consideration is possible in the three-torus case
in section 5. }.
Since the $Q$-exact
structure as well as the $Q$-invariant one is preserved
even after the dimensional reduction,
the dimensionally reduced theory of a cohomological field theory
becomes also a cohomological field theory. Further, the $U(1)$-symmetry
in (\ref{U(1)charge}) remains nonanomalous in the dimensional reduction
to two-dimensions in the
case of the $SU(N)$ gauge group, because it contains no $U(1)$ factor.
Thus, the arguments of the
mass perturbation in section 3.2 can be applied also to the
dimensionally reduced case. Then, the mass perturbation breaks
the ${\cal N}=8$ supersymmetry to ${\cal N}=2,$
and the partition function takes the
form
\begin{equation}
Z_{SU(N)}^{D=2, \;{\cal N}=8}(T^2)=
\sum_{al=N}Z_{SU(l)\otimes {\bf Z}_a}^{D=2, \;{\cal N}=2}(T^2).
\label{D=2N=8partitionfunction1}
\end{equation}
In two-dimensions, by the argument similar as in the four-dimensional
case, the ${\bf Z}_a$ factor of the gauge group yields
$a^{2-1}=a.$ Hence we have
\begin{equation}
Z_{SU(N)}^{D=2,\; {\cal N}=8}(T^2)=
\sum_{al=N}aZ_{SU(l)}^{D=2,\; {\cal N}=2}(T^2).
\label{D=2N=8partitionfunction2}
\end{equation}
Here the ${\cal N}=2$ theory in two dimensions contains Higgs fields,
whose zero momentum modes form continuous spectrum beginning with
zero-energy because
the field space of the Higgs fields is noncompact. This situation
makes ambiguous the relation between the partition function
and the Witten index, so we cannot go along the same line as
in the four-dimensional case.
However, owing to the fact that ${\cal N}=2$ $SU(l)$ SYM theory
in two dimensions is a cohomological field theory,
the partition function
$Z_{SU(l)}^{D=2,\; {\cal N}=2}(T^2)$ can be evaluated.
Let us see it from now.
After an appropriate field redefinition (see Appendix A),
the classical action of the
${\cal N}=2$ SYM theory is written as the BRST exact form
\begin{equation}
S=Q\int d^2\sigma\;{\rm tr}\left(\frac{1}{8g^2}\eta[\phi,\bar{\phi}]
-i\chi\Phi+2g^2\chi H+\frac{1}{2g^2}\psi_{\mu}D_{\mu}\bar{\phi}
\right),
\label{N=2TFT}
\end{equation}
where the BRST transformation is defined by
\begin{equation}
\begin{array}{lll}
QA_{\mu}=\psi_{\mu}, & Q\psi_{\mu}=-iD_{\mu}\phi, & Q\phi=0,\\
Q\chi=H, & QH=[\phi, \chi], & \\
Q\bar{\phi}=\eta,& Q\eta=[\phi, \bar{\phi}]. &
\end{array}
\end{equation}
Note that the $Q$ is nilpotent up to the gauge transformation with the
parameter $\phi,$ which gives a cohomology to equivalent classes with
respect to the gauge transformation. The field contents are as follows.
$A_{\mu}$ is a two-dimensional gauge field, and $\psi_{\mu},$ $\eta,$
$\chi$ together stand for fermions. $\phi$ and $\bar{\phi}$ are
complex Higgs fields, and $H$ is a bosonic auxiliary field.
Ghost number is assigned as $-2$ to $\bar{\phi},$ $-1$ to
$\eta$ and $\chi,$ $0$ to $A_{\mu}$ and $H,$ +1 to $\psi_{\mu},$ +2
to $\phi.$ The ghost number conservation is nonanomalous for the same
reason as in the case of the dimensional reduction of the $U(1)$-symmetry.
The contribution of the path integration of the gauge field
localizes in the configurations determined by
\begin{equation}
\Phi\equiv -2F_{12}=0.
\label{Phiequation}
\end{equation}
The addition of a $Q$-exact term to the action does not
change the theory, if it behaves well at infinity in the field space.
Thus, we may discard the first term in eq. (\ref{N=2TFT})
\begin{equation}
Q\int d^2\sigma\;{\rm tr}\frac{1}{8g^2}\eta[\phi,\bar{\phi}].
\label{firstterm}
\end{equation}
Also, it can be seen that
the partition function is independent of the coupling $g$ for the same
reason.
The localization (\ref{Phiequation})
can be shown by integrating out $H$ and $\chi$ fields in the
$g\rightarrow 0$ limit after the integrals of $\bar{\phi}$ and $\eta.$
Then the partition function becomes
\begin{equation}
\int\frac{{\cal D}A_{\mu}}{{\rm Vol}(SU(l))}
{\cal D}\phi{\cal D}\psi_{\mu}
\;\delta(D_{\mu}D_{\mu}\phi+\{\psi_{\mu},\psi_{\mu}\})
\;\delta(D_{\mu}\psi_{\mu})
\;\delta(D_1\psi_2-D_2\psi_1)\;\delta(F_{12}),
\label{localization}
\end{equation}
which indicates the declared localization.
It should be noted that the localization is determined by the BRST fixed
point $Q\chi=0$ after using the equation of motion of $H.$
Eq. (\ref{localization}) is
not in a suitable form for our purpose, so we will deform the theory
judiciously as in the section 3 in ref. \cite{Witten6}.
We consider the action with the addition of the $Q$-exact term
\begin{equation}
S(t)=S'+tQ\int d^2\sigma\;{\rm tr}\chi\bar{\phi},
\end{equation}
where $S'$ stands for the action $S$ with the term (\ref{firstterm})
eliminated and $t$ is a parameter.
If new BRST fixed points that flow in from infinity when $t$ turns on
do not contribute, the deformed theory coincides with the original one.
After integrating out $H,$ $\eta,$ $\chi$ and $\bar{\phi},$ we end up
with the action in the large $t$ case
\begin{equation}
S(t)= \frac{1}{2g^2t}\int d^2\sigma\;{\rm tr}
[F_{12}(D_{\mu}D_{\mu}\phi+\{\psi_{\mu},\psi_{\mu}\})
-iD_{\mu}\psi_{\mu}(D_1\psi_2-D_2\psi_1)]+O(t^{-2}),
\label{S(t)1}
\end{equation}
which can be again written in the $Q$-exact form
\begin{equation}
S(t)=Q\left[\frac{i}{2g^2t} \int d^2\sigma\;{\rm tr}
F_{12}D_{\mu}\psi_{\mu}+O(t^{-2})\right].
\label{S(t)2}
\end{equation}
Considering the case $t=-iu$ with $u$ large real positive,
the $\phi$-integration yields $\delta (D_{\mu}D_{\mu}F_{12}).$
Using the normalizability of $F_{12},$ it means that the localization
realizes at the solutions of $D_{\mu}F_{12}=0,$ which contain
extra components adding to the localization
point of the original theory $F_{12}=0.$ Arising of the extra components
is a signal of the flow of new BRST fixed points from infinity.
In fact, in the above process, the contribution of the $\chi$ and
$\bar{\phi}$ integrals localizes the points
\begin{eqnarray*}
\chi & = & \frac{1}{2t}D_{\mu}\psi_{\mu}, \\
\bar{\phi} & = & -\frac 1t F_{12}+\frac{1}{2t^2}(D_{\mu}D_{\mu}\phi
+\{\psi_{\mu},\psi_{\mu}\}),
\end{eqnarray*}
which appear first when $t$ turns on.
So they are the new fixed points flowing in from infinity.
Thus in general,
the deformed cohomological field theory does not equivalent
to the original one. However, there is a possibility that the
BRST invariant operators with the following feature exist ---
in calculation of their correlators the extra components do not
contribute. If there are such operators, the deformed theory coincides
to the original one with respect to the restricted set of the operators.
Indeed, we can find such operators. For the following BRST invariant
operators
\begin{eqnarray}
& & \omega\equiv \int d^2\sigma \; {\rm tr} \;(-i\phi F_{12}+\psi_1\psi_2), \\
& & \beta (\phi) = {\rm tr}\;(\;{\rm polynomial\; of \;}\phi), \nonumber
\end{eqnarray}
we consider the unnormalized expectation value in the deformed theory
\begin{equation}
\left\langle e^{\omega}\beta(\phi)\right\rangle'\equiv
\int\frac{{\cal D}A_{\mu}}{{\rm Vol}(SU(l))}{\cal D}\phi{\cal D}\psi_{\mu}
\;\beta(\phi)\;e^{-S(-iu)+\omega}.
\label{expectationvalue}
\end{equation}
Here due to the $e^{\omega}$ factor, we can take the limit $u=\infty$
without changing the behavior of the fields at the infinity,
and thus the limit
does not change the value of (\ref{expectationvalue}).
Further integrating out $\phi$ and $\psi_{\mu},$ we end up with
\begin{equation}
\left\langle e^{\omega}\beta(\phi)\right\rangle'=
\int\frac{{\cal D}A_{\mu}}{{\rm Vol}(SU(l))}\;
\beta\left(-i\frac{\delta}{\delta F_{12}}\right)\;\delta(F_{12}),
\label{expectationvalue2}
\end{equation}
where the factor $\delta(F_{12})$ (with a finite degree of the
derivative $\frac{\delta}{\delta F_{12}}$)
indicates no contributions of the
extra components. Therefore, $\left\langle e^{\omega}\beta(\phi)\right\rangle'$
coincides with the unnormalized expectation value in the original theory
which we denote by $\left\langle e^{\omega}\beta(\phi)\right\rangle.$
Now we can manage to compute the partition function
$Z_{SU(l)}^{D=2,\; {\cal N}=2}(T^2)=\langle 1\rangle.$ Notice
that $\omega$ has the ghost number $+2,$ and thus from the ghost number
conservation we can show
$\langle 1\rangle=\langle e^{\omega}\rangle.$ This is equal to the $\beta=1$ case
of eq. (\ref{expectationvalue2}), so we find
\begin{equation}
Z_{SU(l)}^{D=2,\; {\cal N}=2}(T^2)=\int
\frac{{\cal D}A_{\mu}}{{\rm Vol}(SU(l))}\;
\delta(F_{12}),
\label{partitionfunction3}
\end{equation}
which counts the number of the small gauge inequivalent configurations
satisfying $F_{12}=0.$ The number is unity,
since the two-dimensional $SU(l)$ gauge theory
has no nontrivial winding number.
In the consequence we have
\begin{equation}
Z_{SU(l)}^{D=2,\; {\cal N}=2}(T^2)=1.
\end{equation}
It can be confirmed also
by doing the concrete calculation, for example, employing the $A_1=0$
gauge fixing in eq. (\ref{partitionfunction3}).
Plugging this into eq.(\ref{D=2N=8partitionfunction2}) we get the
result
\begin{equation}
Z_{SU(N)}^{D=2,\; {\cal N}=8}(T^2)=\sum_{al=N}a.
\end{equation}
Finally by taking account into the ${\bf Z}_N$ factor
as in the four dimensional case, the second factor
$Z_{SU(N)/{\bf Z}_N}^{D=2,\;{\cal N}=8}(T^2)$ of eq. (\ref{ZIIAMS})
is obtained as
\begin{equation}
Z_{SU(N)/{\bf Z}_N}^{D=2,\;{\cal N}=8}(T^2)=
\frac 1N Z_{SU(N)}^{D=2,\; {\cal N}=8}(T^2)
=\sum_{a|N}\frac 1a.
\label{resultsecondfactor}
\end{equation}
\subsection{Result of IIA Matrix String Partition Function}
Now we can write down the result of the partition function of
the IIA matrix string theory. Substituting eqs. (\ref{firstfactor})
and (\ref{resultsecondfactor}) into eq. (\ref{ZIIAMS}), we have
\begin{equation}
Z_{{\rm IIA-MS}}=\left(\sum_{a|N}\frac 1a\right)\sum_{p\in{\bf Z}}
e^{-\frac{RT}{2}Ng^2p^2},
\label{IIAresult1}
\end{equation}
which coincides the result obtained in the strongly coupled limit
($g^2\rightarrow \infty$) by Kostov and Vanhove \cite{Kostov-Vanhove}.
They have conjectured that their result holds irrespectively of the
strength of the coupling, and they called this property {\it exact
quasi classics.} Since our calculation has been exactly performed
without any approximation, it gives a proof of the {\it exact
quasi classics.}
Recalling the relations (\ref{RLrelation}) and (\ref{IIAggs}),
we rewrite the result (\ref{IIAresult1})
in variables in string theory
\begin{equation}
Z_{{\rm IIA-MS}}=\left(\sum_{a|N}\frac 1a\right)\sum_{p\in{\bf Z}}
e^{-T\frac{NL}{4\pi}\frac{p^2}{(g_sl_s)^2}}.
\label{IIAresult2}
\end{equation}
There are two comments in order.
First, the second factor in (\ref{IIAresult2})
represents a certain nonperturbative effect, which corresponds to
creation and annihilation of D-particle and anti-D-particle pairs.
Such phenomena as creation/annihilation of D- and anti-D- objects
cannot be seen in the M(atrix) theory,
because in the infinite momentum
frame anti-D-particles
in the M(atrix) theory are integrated out and do not appear.
It can be seen first after compactified to the IIA matrix string theory.
Second, there is no perturbative correction in the formula
(\ref{IIAresult2}). It agrees to nonrenormalization theorems in
perturbative superstring theory by Martinec \cite{Martinec},
which tells that the 0-, 1-, 2-, and 3-point functions of massless
string vertex operators receive no perturbative corrections in
flat ten-dimensional backgrounds.
\section{IIB Matrix String Partition Function}
Here, we compute the partition function of IIB matrix string
theory in the ten-dimensional IIB limit, and compare the exact result
of the IKKT model by
Moore-Nekrasov-Shatashvili \cite{Moore-Nekrasov-Shatashvili}.
We consider the three-dimensional SYM theory with periodic boundary
condition as the IIB matrix string theory for the same reason
as in the
IIA case.
With respect to the gauge group, the $U(1)$ part of the gauge group
$U(N)=U(1)\times (SU(N)/{\bf Z}_N)$
for the field contents $(A_{\mu}, X, \theta)$ corresponds to
the center of mass coordinates of the strings in transverse
directions in the ten-dimensional IIB limit ``$R_1R_2\rightarrow\infty$
with $g_s$ fixed,'' where the $U(1)$ part of the gauge field together
with that of $X$ become the transverse coordinates. Thus we fix
the $U(1)$ part of the gauge field as well as the Higgs fields,
and consider the partition function of the
three-dimensional ${\cal N}=8,$ $SU(N)/{\bf Z}_N$ SYM theory as that of
the IIB matrix string theory
\begin{equation}
Z_{{\rm IIB-MS}}=Z_{SU(N)/{\bf Z}_N}^{D=3, \; {\cal N}=8}(T^3).
\label{ZIIBMS}
\end{equation}
The three-torus, where the SYM theory is defined on, is taken
to be rectangular of the size $T\times R_1 \times R_2,$ with the
Euclidean signature.
(We performed the Wick rotation as in the IIA case.)
In this case, we can also use the dimensionally reduced version of
the arguments of the mass perturbation in the four dimensions,
due to the following two reasons. One is that the dimensional reduction
of a cohomological field theory is also a cohomological field theory.
The other
is that in odd dimensions
the dimensionally reduced version of the $U(1)$-symmetry
in eq. (\ref{U(1)charge}) is anomaly free.
Going on along the same line as in the IIA case, we have
\begin{eqnarray}
Z_{SU(N)}^{D=3,\; {\cal N}=8}(T^3) & = &
\sum_{al=N}Z_{SU(l)\otimes{\bf Z}_a}^{D=3, \; {\cal N}=2}(T^3) \nonumber \\
& = & \sum_{al=N}a^2Z_{SU(l)}^{D=3, \; {\cal N}=2}(T^3).
\label{ZD=3N=8SYM}
\end{eqnarray}
Here, there exists a Higgs field in the three-dimensional ${\cal N}=2$
theory. For the same reason as before, we cannot relate
the partition function
directly to the Witten index. However, if considering
the ten-dimensional IIB limit (\ref{10DIIBlimit})
which in fact we are interested in, we can proceed further.
Note that in the large volume limit physics becomes independent of
the boundary condition. So we can evaluate
$Z_{SU(l)}^{D=3, \; {\cal N}=2}(T^3)$
by adopting a twisted boundary condition instead of the periodic
boundary condition.
\subsection{${\cal N}=2$ partition function with a twisted boundary
condition}
Here we consider the three-dimensional ${\cal N}=2$ $SU(l)$ SYM theory
with the twisted boundary condition
\begin{eqnarray}
A_i(t,\sigma_1,\sigma_2) & = & A_i(t+T, \sigma_1,\sigma_2) \nonumber \\
& = & PA_i(t,\sigma_1+R_1, \sigma_2)P^{-1} \nonumber \\
& = & QA_i(t,\sigma_1, \sigma_2+R_2)Q^{-1},
\label{twistedBC}
\end{eqnarray}
where $i=1,2,$ we took the $A_0=0$ gauge fixing, and $P$ and $Q$ are
$SU(l)$ matrices satisfying $PQ=QPe^{2\pi i/l}.$ For example, $P$
and $Q$ can be represented as
\begin{equation}
P=e^{-\pi i\frac{l+1}{l}}\left[\begin{array}{ccccc}
0 & 1 & & & \\
& 0 & 1 & & \\
& &\ddots & \ddots & \\
& & & \ddots & 1 \\
1 & & & & 0
\end{array}\right],
\hspace{1cm}
Q=e^{-\pi i\frac{l-1}{l}}\left[\begin{array}{ccccc}
1 & & & & \\
& e^{2\pi i/l}& & & \\
& & e^{4\pi i/l} & & \\
& & & \ddots& \\
& & & & e^{2\pi i(l-1)/l}
\end{array}\right].
\end{equation}
Supersymmetry requires that the other fields (Higgs $\phi$ and complex
fermion $\lambda$) satisfy the same boundary condition
as eq. (\ref{twistedBC}). Under this
boundary condition, zero momentum modes become trivial
\begin{equation}
A_i^{(0)}=\phi^{(0)}=\lambda^{(0)}=0,
\label{****}
\end{equation}
because the constant traceless hermitian matrices
commuting with
$P$ and $Q$ simultaneously do not exist
except the trivial case (\ref{****}).
In this situation, spectra appearing in the theory are discrete,
and thus the partition function coincides the Witten index.
Let us consider the Witten index. The argument below is
a three-dimensional analogue of Witten's consideration
in the four-dimensional case \cite{Witten6}. Also, though it is
briefly reported in \cite{Affleck-Harvey-Witten}, we will discuss it
in order to make this paper more self-contained.
Our problem is now
reduced to counting the number of vacua, i.e. the number
of (small) gauge inequivalent classes of the classical zero energy
states.
The gauge transformation preserving both of the boundary condition
(\ref{twistedBC}) and the gauge condition $A_0=0$ is generated by
the time-independent
$SU(l)$ matrix of the following boundary condition
\begin{eqnarray}
U(\sigma_1,\sigma_2) & = &
e^{2\pi ik_1/l}PU(\sigma_1+R_1,\sigma_2)P^{-1} \nonumber \\
& = & e^{2\pi ik_2/l}QU(\sigma_1,\sigma_2+R_2)Q^{-1}
\label{UBC}
\end{eqnarray}
where $k_i=0,1,\cdots,l-1.$
The classical zero energy state is given by the configuration
$$
A_i=-i(\partial_iU)U^{-1}
$$
with the other fields nil. Here, if this $U$ can be continuously deformed
to the identity, there are no nontrivial sectors, and thus the vacuum
is unique modulo small gauge transformations. We will see that
it is in fact so. The $U$ can be written by the $SU(l)$ matrix
$\tilde{U}$ satisfying the simpler boundary condition
\begin{equation}
U=(Q^{-1})^{k_1}P^{k_2}\tilde{U},
\label{UUtilde}
\end{equation}
where
\begin{eqnarray}
\tilde{U}(\sigma_1,\sigma_2) & = &
P\tilde{U}(\sigma_1+R_1,\sigma_2)P^{-1} \nonumber \\
& = & Q\tilde{U}(\sigma_1,\sigma_2+R_2)Q^{-1}.
\label{UtildeBC}
\end{eqnarray}
We consider a topological classification of gauge transformations
with the boundary condition (\ref{UtildeBC}).
Because of $\pi_0(SU(l))=0,$
by a suitable continuous translation in a group manifold of $SU(l),$
we can always
start with $\tilde{U}(0,0)=1.$ Then, using eq. (\ref{UtildeBC})
we see that
$$
\tilde{U}(0,0)=\tilde{U}(R_1,0)=\tilde{U}(0,R_2)=\tilde{U}(R_1,R_2)=1,
$$
i.e. $\tilde{U}$'s on the vertices of a square with the size
$R_1\times R_2$ in $(\sigma_1, \sigma_2)$-space are all identity. From
the fact $\pi_1(SU(l))=0,$ $\tilde{U}$ on the edges of the square can
be taken identity by a continuous deformation. Further, using
$\pi_2(SU(l))=0,$ we continuously deform to $\tilde{U}=1$ everywhere
in the square.
Using $\pi_0(SU(l))=0$ again, by a continuous translation
the $U$ in eq. (\ref{UUtilde}) can be taken identity
on the square.
Thus, it is confirmed that there is no nontrivial
topological vacuum sector. We conclude that the Witten index is unity,
which leads to
\begin{equation}
Z_{SU(l)}^{D=3, \; {\cal N}=2}(T^3)|_{{\rm twisted \;B.C.}}=1.
\label{ZtwistedBC}
\end{equation}
\subsection{Result of IIB Matrix String Partition Function}
As discussed before, in the ten-dimensional IIB limit
we can replace the value of the partition function
$Z_{SU(l)}^{D=3, \; {\cal N}=2}(T^3)$ with that of eq. (\ref{ZtwistedBC})
\begin{equation}
Z_{SU(l)}^{D=3, \; {\cal N}=2}(T^3)=1.
\label{ZperiodicBC}
\end{equation}
Substituting this into eq. (\ref{ZD=3N=8SYM}), we have
\begin{equation}
Z_{SU(N)}^{D=3,\; {\cal N}=8}(T^3) = \sum_{al=N}a^2,
\end{equation}
and thus the partition function of the IIB matrix string theory is
obtained as
\begin{eqnarray}
Z_{{\rm IIB-MS}} & = & Z_{SU(N)/{\bf Z}_N}^{D=3, \; {\cal N}=8}(T^3) \nonumber \\
& = & \frac{1}{N^2}Z_{SU(N)}^{D=3,\; {\cal N}=8}(T^3) \nonumber \\
& = & \sum_{a|N}\frac{1}{a^2},
\label{ZIIBMS2}
\end{eqnarray}
which is valid in the ten-dimensional IIB limit (\ref{10DIIBlimit}).
We should remark that the result (\ref{ZIIBMS2})
coincides the result of the IKKT model \cite{IKKT} by Moore, Nekrasov
and Shatashvili \cite{Moore-Nekrasov-Shatashvili}.
It is quite nontrivial and might be a signal of the equivalence
between the two IIB matrix models arising from the different roots
--- one is a compactification of the BFSS M(atrix) theory,
and the other is a
matrix regularization of the worldsheet action of type IIB superstring
in Schild gauge.
Also, the result (\ref{ZIIBMS2}) has neither perturbative nor
nonperturbative correction. With respect to the former it agrees again
with the nonrenormalization theorems in perturbative superstring theory
\cite{Martinec}.
\section{Conclusions}
We have considered the IIA and IIB matrix string theories derived from
the M(atrix) theory via toroidal compactification. We have summarized
that string interactions emerge as the Yang-Mills instantons and
have shown
that in the IIB matrix string theory the chiral spinors are correctly
reproduced.
As a preparation for computation of
the matrix string partition functions, we have calculated the partition
function of four-dimensional SYM theory on $T^4,$ by mapping
the theory to a cohomological field theory.
Here, considering the mass perturbation (\ref{masspert}) has made
the calculation easier. We have shown the invariance of the partition
function under the mass perturbation in a different fashion from
the argument by Vafa and Witten \cite{Vafa-Witten}.
We have exactly computed the partition function of the IIA matrix
string theory by mapping the theory into a cohomological field theory.
Our result coincides with the result obtained in the infra-red limit
by Kostov and Vanhove \cite{Kostov-Vanhove}, and thus gives a proof of
the {\it exact quasi classics} of the theory conjectured by them.
The formula for the partition function receives no perturbative
correction, which is in conformity with nonrenormalization theorems
in perturbative superstring theory by Martinec \cite{Martinec}.
Also, there exist some nonperturbative corrections
which come from the $U(1)$ electric flux in the SYM theory
and which are interpreted
as creation and annihilation of D-particle and anti-D-particle pairs.
Such phenomena have been reported in high energy scattering of strings
in ref. \cite{GHV}. It may be interesting to deepen the meaning of our
result from the line of the high energy scattering.
Further, we have evaluated the partition function of the IIB matrix
string theory in the ten-dimensional IIB limit by a similar method as
in the IIA case.
Our result receives neither perturbative and nonperturbative
corrections, which with respect to the former agrees with the
nonrenormalization theorems again.
Also, our result coincides with the exact result of
the partition function of the IKKT model by
Moore, Nekrasov and Shatashvili \cite{Moore-Nekrasov-Shatashvili}.
Although both of the IIB matrix string theory
and the IKKT model are considered
to give type IIB string theory,
they have arised from the distinctive
origins, and the relation between them have not been
clarified yet. Thus our result is quite nontrivial, and may suggest
the equivalence of those two models.
In the IIB matrix string theory, the S-duality is well understood from
a geometry of the two-torus, but the ten-dimensional Lorentz symmetry
is not manifest. On the other hand, in the IKKT model side,
while there is a
manifest Lorentz symmetry, we have not been able to see the S-dual
structure.
In this situation, it seems to be an important step to explore the
equivalence and to establish the precise correspondence between them
toward constructing the nonperturbative definition which manifestly
realizes both of
the Lorentz symmetry and the S-dual structure.
\vspace{3cm}
\begin{large}
{\bf Acknowledgements}
\end{large}
\vspace{7mm}
The preliminary version of this work was presented
in KEK theory workshop '99. The author would like to thank the
organizers and participants of the workshop, and especially
N. Ishibashi, S. Iso, H. Kawai, T. Kuroki, Y. Okada, K. Okuyama
and A. Tsuchiya for valuable conversations and encouragements.
The research of the author is supported by the Japan Society for
the Promotion of Science under the Postdoctral Research Program.
\vspace{3cm}
\begin{large}
{\bf Note Added}
\end{large}
\vspace{7mm}
While writing up the manuscript, I received the papers \cite{GS,P}
which discuss issues close to this work.
In \cite{GS}
the thermodynamic partition function of the IIA matrix
string theory is calculated in the $g_s\rightarrow 0$ limit,
and in \cite{P} some thermodynamic properties of
the IIA and IIB matrix string theories are discussed.
\newpage
\begin{large}
{\bf Appendix}
\end{large}
|
\section*{Acknowledgements}
We warmly acknowledge W.~Vogelsang for providing us with the
NLO-evolved parton densities in the photon, and G.~Ridolfi for
discussions. This work was supported in part by the EU Fourth
Framework Programme `Training and Mobility of Researchers', Network
`Quantum Chromodynamics and the Deep Structure of Elementary
Particles', contract FMRX-CT98-0194 (DG 12-MIHT).
|
\section{INTRODUCTION}
The concept of coherent states (CS) is widely used in different
fields of physics and mathematics (see for example Refs.
\cite{Per} - \cite{KS}). In particularly, it plays an important
role in the
Berezin quantization scheme \cite{Ber}, in the analysis of
growth
of functions holomorphic in a complex domain \cite{Vourd}, in a
general theory of phase space quasiprobability distributions
\cite{BM}, and in a quantum state engineering \cite{Vo}.
It is necessary to note that in present no a unified definition
of such states exists in the literature and different authors
mean different things when speaking about them. Nevertheless, a
careful analysis (see for example Ref. \cite{Klaud}) shows that
almost all definitions
have some common points that can be taken as a general
definition. Following Klauder \cite{Klaud} I mean by coherent
states such states that satisfy the following defining
properties:
(1) CS are defined by vectors $\psi _z(x,t)$ which belong to a
Hilbert space $H$ of the states of a quantum system
with scalar product $\langle \cdot |\cdot \rangle $;
(2) The parameter $z$ takes continuous values from a domain
${\cal D}$ of an $n$-dimensional complex space;
(3) There exists a measure $\mu =\mu (z,\bar z)$ (the bar over
a symbol indicates complex conjugation) that realizes the
resolution of the identity operator $\Bbb I$ acting in $H$
in terms of the projectors on the vectors $\psi _z$
\begin{equation}
\int _{\cal D} d\mu |\psi _z\rangle \langle \psi _z|=
\Bbb I\,;
\label{Id}
\end{equation}
(4)CS have to prove the property of a temporal
stability. By temporal stability I mean that the vectors
$\psi _z(x,t)$ remain coherent at all times i.e. satisfy the
properties 1.-3. at all times.
To satisfy this condition I shall
assume that the functions $\psi _z(x,t)$ are solutions of the
Schr\"o\-din\-ger equation
\[(i\partial _t-h_0)\psi _z(x,t) =0 \]
where $h_0$ is the Hamiltonian of a given quantum system which
in general can depend on time. Operator $h_0$ is supposed to be
Hermitian in $H$ and to have a unique self-adjoint extension.
The Eq. (\ref{Id}) should be understood in a weak sense.
This means that it is equivalent to the following relation
\[
\int _{\cal D} d\mu \langle \psi _a|\psi _z\rangle
\langle \psi _z|\psi _b \rangle =
\langle \psi _a|\psi _b\rangle
\]
which should hold for all $\psi _{a,b}$ from a dense set in
$H$.
Transparent potentials have many remarkable properties.
For instance, a quantum particle prove no reflection in the
scattering process on such a potential. Another remarkable
property is that each level in the discrete spectrum
of such a potential occupies a
preassigned position, which is controlled by values of the
parameters the potential depend on. The
discrete spectrum levels may even be
situated in the middle of the continuous spectrum. In the
latter case we have {\it completely transparent potentials}
\cite{Stahlh}.
Transparent potentials find a more significant application in
soliton theory. There is a marvelous vast literature on
this subject. I cite here only a monograph \cite{MatvSal}.
Because of their remarkable properties transparent potentials
would find an application in pseudopotential theories. Recently
they have been used to describe relaxation processes in Fermi
liquid \cite{VVT}.
CS for transparent potentials are very far from being explored.
It may be explained by the fact that up to now no systematic
way is
known for their investigation.
No a clear algebraic structure related to these potentials is
known and therefore well known algebraic methods \cite{Per}
prove
to be a little suitable in this context. No simple ladder
operators for the discrete spectrum eigenfunctions of
transparent potentials are known as well and therefor we can
not use the approach of Ref. \cite{MM} for this purpose. An
approach based on the uncertainty relation \cite{NiSai0} is not
consistent with the property 3. mentioned above and therefor it
should be rejected.
A conjecture has been advanced recently \cite{BSjetp,NF} to use
Darboux
transformation operator approach for investigating the CS of
those system that is related by Darboux transformation
with that for which the CS are known.
Let us have an exactly
soluble Hamiltonian $h_0=-\partial _x^2 +V_0(x,t)$ for which
the CS $\psi _z(x,t)$ are known and we want to obtain the CS
for another Hamiltonian $h_1=-\partial _x^2 +V_1(x,t)$
related with $h_0$ by the Darboux transformation operator that
I shall denote by $L$. In general it should be a {\it
nonstationary} Darboux transformation operator defined by the
following intertwining relation \cite{BSPL}
\[L(i\partial _t-h_0)=(i\partial _t-h_1)L\, .\]
If such an operator $L$ is known then solutions of the
transformed Schr\"o\-din\-ger equation
(i.e. the Schr\"o\-din\-ger equation with the
Hamiltonian
$h_1$) can easily be obtained by the action of the operator $L$
on solutions of the initial Schr\"o\-din\-ger equation
(i.e. the Schr\"o\-din\-ger equation
with
the Hamiltonian $h_0$). It is clear that the functions
$\varphi _z(x,t) =L\psi _z(x,t)$ will satisfy all the
properties of the
CS enumerated above except may be for the property 3. One of
the main goal of this paper is to prove that in the case of
soliton potentials this property is fulfilled.
I would like to mention that this approach has been
successfully
applied to study CS of anharmonic oscillator Hamiltonians with
equidistant and quasiequidistant spectra \cite{BSjetp} and CS
of
nonstationary soliton potentials \cite{jetp98E} that are
related with
soliton solutions of the Kadomtsev-Petviashvili equation. With
the help of this approach a classical counterpart of the
Darboux
transformation has been formulated and shown that at classical
level this transformation leads to a distortion of a phase
space \cite{SamJMP}. CS of a one-soliton potential have been
investigated as well and their supercoherent structure has been
revealed \cite{NF}. In this paper I generalize these results to
a multisoliton case.
The paper is organized as follows.
In the Section 2 I give well-known results for CS of the free
particle in preparation for their application in the following
sections. In the Section 3 the Darboux
transformation operator from the solutions of the
Schr\"o\-din\-ger equation
with zero potential to the solutions of the same equation with
solitons potential is analyzed as an operator acting in the
Hilbert space of the states of a free particle. It is shown
that it
can not realize a mapping of Hilbert spaces since it is not
defined in the whole Hilbert space and can not be extended to
the whole Hilbert space. Isomeric operators expressed in terms
of continuous bases similar to these previously proposed by
L.D. ~Faddeev \cite{Fadd} and
analyzed by D.L. ~Pursey \cite{Pur} for the case of purely
discrete basis sets are introduced. These operators realize
a polar decomposition of Darboux transformation operators. A
quazispectral representation of the Darboux transformation
operator and
its inverse in terms of continuous bases are obtained. In the
Section 4. different systems of CS are introduced for soliton
potentials. It is established that the resolution of the
identity
operator exists in every case. Explicit expressions for
measures that realize this equality are found.
A brief conclusion brings a paper to a close.
\section{FREE PARTICLE COHERENT STATES}
In this section I give a brief overview of well-known
properties of the Hilbert space of states a free particle
(see Ref. \cite{Mil} and references therein) and corresponding
CS \cite{MM} we need for subsequent analysis.
Annihilation $a$ and creation $a^+$ operators
\[a=(i-t)\partial _x+ix/2,\quad a^+=(i+t)\partial _x-ix/2\]
form the Heisenberg-Weil subalgebra of the six-dimensional
Schr\"o\-din\-ger
algebra which is a symmetry algebra of the equation with
zero potential. Solutions of the free particle
Schr\"o\-din\-ger equation
which are square integrable over full real axis
$\Bbb R =(-\infty , +\infty )$ with respect to the Lebesgue
measure are the eigenstates of the symmetry operator
$K_0=aa^++a^+a$, $K_0\psi _n(x,t) =(2n+1)\psi _n(x,t) $. Their
coordinate
representation is as follows
$$
\begin{array}{rl}
\psi _n(x,t) =
&
(-i)^n(n!2^n\sqrt {2\pi })^{-1/2}
(1+it)^{-1/2} \\
\times &
\exp \!
\left[-in\arctan t+{\textstyle \frac {y^2}2}(it-1)\right]
H_n\left( y \right)\, ,\\
y=
&
{\displaystyle \frac{x}{\sqrt{2+2t^2}}}\, .\vphantom{\left)
{\displaystyle \frac AB} \right)}
\end{array}
$$
Operators $a$ and $a^+$ are the ladder operators for the basis
functions $\psi _n$:
$a\psi _n=\sqrt n\psi _{n-1}$,
$a^+\psi _n=\sqrt{n+1}\psi _{n+1}$, and $a\psi _0=0$.
By ${\cal L}_0 $ I denote the lineal of the functions $\psi_n
$,
$n=0,1,\ldots $ which is the space of all finite linear
combinations of the functions $\psi_n $ with the coefficients
from the field $\Bbb C$. The operators $a$ and $a^+$ being
linear are defined for all elements from ${\cal L}_0 $ and
${\cal L}_0 $ is
invariant with respect to the action of these operators.
Since the momentum operator $p_x=-i\partial _x$ and the initial
Hamiltonian $h_0$ are expressed in terms of $a$ and $a^+$:
$p_x=-(a+a^+)/2$, $h_0=p_x^2$, these operators are defined in
${\cal L}_0 $ and map this space into itself.
The Hilbert space of the states of the free particle, $H$, is
defined as a closure of the lineal ${\cal L}_0 $ with respect
to the
measure generated by the scalar product
$\langle \psi _a|\psi _b\rangle $, $\psi _{a,b}\in {\cal L}_0
$, which
is defined by the Lebesgue integral.
The functions $\psi _n$ form an orthonormal basis in $H$,
$\langle \psi _n|\psi _k\rangle =\delta _{nk}$.
It is well-known \cite{Smirn,RS}
that the operators $p_x$ and $h_0$ initially defined on ${\cal
L}_0 $
have unique self-adjoint extensions and consequently they
are essentially self-adjoint in $H$.
The spectrum of $h_0$ and $p_x$ is purely continuous. They have
common eigenfunctions $\psi _p=\psi _p(x,t)$:
$p_x\psi _p=p\psi _p$, $h_0\psi _p=p^2\psi _p$, $p\in \Bbb R$,
which do not belong to $H$ but belong to a more wide space
$H_-$ of the linear functionals over $H_+$, $H_+\subset
H\subset H_-$
(so called Gelfand triplet). We can choose the Hilbert-Schmidt
equipment of the space $H$ by letting $H_+=K_0^{-1}H$ since
$K_0^{-1}$ is a Hilbert-Schmidt operator. We refer a reader
to Refs. \cite{GSh,GV,BerezShub} for more details on the nested
Hilbert space.
The coordinate representation of the functions $\psi _p(x,t)$
is
well-known and I omit it here.
The functions $\psi _p$ form an orthonormal and complete (in
the sense of generalized functions) basis in $H$,
$\langle \psi _p|\psi _q\rangle =\delta (p-q)$. The
completeness condition is expressed symbolically in terms of
the projectors onto these functions as follows
\begin{equation}
\int dp|\psi _p\rangle \langle \psi _p |=\Bbb I\, .
\label{PI}
\end{equation}
I do not indicate the limits of integration in the integrals
along the whole real axis. This equality should be understood
in a weak sense. This means that it is equivalent to
\[
\int dp\langle \psi _j|\psi _p\rangle
\langle \psi _p |\psi _k\rangle =\delta _{jk}\, , \quad
j,k=0,1,\ldots
\]
where $\psi _k$, $k=0,1,\ldots $ are orthonormal basis
functions in the space $H$.
The free particle CS may be obtained by applying a displacement
operator in the Heisenberg-Weil group to the vacuum vector
$\psi _0$:
\[
\psi _z(x,t) =\exp (za^+-\bar z a)\psi _0(x,t)\,, \quad z\in
\Bbb C\,.
\]
These vectors are also the eigenvectors of the annihilation
operator $a\psi _z =z\psi _z$. The vectors $\psi _z\in H$
belong to a more wide set then ${\cal L}_0 $. Their Fourier
decomposition in terms of the basis $\psi_n$ has the form
\begin{equation}
\begin{array}{rl}
\psi _z=
&
\Phi \sum \nolimits _n a_n z^n \psi _n\,, \label {psz}
\end{array}
\end{equation}
\vspace{-4ex}
\[
\begin{array}{rl}
\Phi =
&
\Phi (z,\bar z)=\exp (-z\bar z /2)\,, \\
a_n=
&
(n!)^{-1/2},\quad z\in \Bbb C\,.
\end{array}
\]
The vectors $\psi _z(x,t) $ satisfy all the properties
enumerated in
the Introduction. In particular, the measure
$d\mu =d\mu (z,\bar z)$ from the relation (\ref{Id}) is well-%
known: $d\mu =dxdy/\pi $, $z=x+iy$ and the domain of
integration ${\cal D}$ is the whole complex plane $\Bbb C$. In
what follows I will not indicate the domain of integration in
the integrals over the measures. Integration will be always
extended over the whole complex plane. Finally I give a
coordinate representation of the free particle CS
$$
\begin{array}{rl}
\psi _z(x,t) =&
\hspace{-4pt}
(2\pi )^{-1/4}(1+it)^{-1/2}\\
&
\hspace{-14pt}
\times \exp \left[ -{\textstyle \frac 14}(z+\bar z)^2+
{\displaystyle \frac{(x+2iz)^2(it-1)}{4(1+t^2)}} \right] .
\end{array}
$$
I use the notation $x$ as the spatial coordinate and as the
real part of a complex number $z$. I hope that it will not cause
a confusion since these quantities will never appear in the
same formula.
\section{DARBOUX TRANSFORMATIONS AND ISOMETRIC OPERATORS}
In this section I give an analysis of Darboux transformation
operator $L$ as an operator defined in the Hilbert space $H$.
I would like to stress that this operator is unbounded and can
not be defined over the whole space $H$. It has a domain of
definition which is a subset of $H$ and will be specified.
Moreover, it domain of values does not coincide with $H$.
Therefor this operator can not realize shifting between Hilbert
spaces
contrary to published assertion \cite{Montem}.
To obtain $N$-soliton potential I use the Darboux
transformation
operator approach elaborated in details in Ref. \cite{MatvSal}.
The
action of this operator on a sufficiently smooth function is
defined by the formula
\[L\psi =W^{-1}(u_1,\ldots ,u_N) W(u_1,\ldots ,u_N,\psi )\]
where $W$ stands for the usual symbol of a Wronskian. In the
case when the initial potential $V_0$ does not depend on time,
the functions $u_k=u_k(x,t)$ being solutions of the initial
Schr\"odinger equation may be eigenfunctions of the initial
Hamiltonian as well
$h_0u_k=a _ku_k$ and in general are not supposed to satisfy
any boundary conditions. In this case the transformation
operator $L$ does not depend on time and transforms solutions
of the initial Schr\"o\-din\-ger equation onto solutions
of the
Schr\"o\-din\-ger equation with the potential
\[V_1=V_0-2\partial _x^2\log W(u_1,\ldots ,u_N)\]
which is independent on time. In this paper we need not to use
time dependent Darboux transformation which was proposed by
V. ~Matveev and M. ~Salle (see Ref. \cite{MatvSal}) and
advanced by V. ~Bagrov and B. ~Samsonov \cite{Rev}.
To obtain an $N$-soliton potential we should take $V_0=0$ and
specify the transformation functions $u_k$ as follows
\cite{MatvSal}:
\[
\begin{array}{rl}
u_{2k-1}=
&
\cosh (a_{2k-1}x+b_{2k-1})\,,\\
u_{2k}=
&
\sinh (a_{2k}x+b_{2k})\,,\\
h_0u_k=
&
-a_k^2u_k\,, \ k=1,2,\ldots ,N\,,\\
&
a_1<a_2<\ldots <a_N.
\end{array}
\]
The time dependent phase factors are omitted from these
functions since they do not affect all the results.
In general the Wronsky determinant contains $N!$ summands. I
would like to stress that in a special case of soliton
potentials this determinant may be substantially simplified and
presented as a sum of $2^{N-1}$ hyperbolic cosines \cite{CSh}
\[
W(u_1,\ldots ,u_N)=2^{1-N}\!\!\!\sum_{(\varepsilon_1,\ldots
,\varepsilon_N)}^{2^{N-1}}
\!\!\!\varepsilon_2\varepsilon_4\cdots\varepsilon_{p}\]
\vspace{-3ex}
\[\times \prod_{j>i}^{N}(
\varepsilon_{j}a_j-
\varepsilon_{i}a_i)\cosh [\sum_{l=1}^N\varepsilon_l
(a_{l}x+b_l)]\,,
\]
where $\varepsilon_i=\pm 1$, the value of the subscript $p$ at
$\varepsilon _p$ should be taken equal to $N$ for even $N$
values and to $N-1$ for odd $N$ values, the summation runs over
all ordered and nonidentical sets
$(\varepsilon_1,\ldots ,\varepsilon_N)$
(the sets $(\varepsilon_1, \ldots ,\varepsilon_N)$ and
$(-\varepsilon_1, \ldots ,-\varepsilon_N)$ are declared to be
identical).
It can be shown \cite{MatvSal} that the potential so obtained
is
regular and bounded from below. This implies that the
Hamiltonian $h_1=-\partial _x^2+V_1$ is essentially self-%
adjoint in $H$. It has a mixed spectrum. The position of the
discrete spectrum levels is defined by the values of the
parameters $a_k$: $E_k=-a_k^2$. Corresponding eigenfunctions
have the form \cite{BSTMF}
\[
\begin{array}{rl}
\varphi _k=
&
N_kW^{(k)}(u_1,\ldots ,u_N)/W(u_1,\ldots ,u_N)\,,\\
N_k=&
({\textstyle \frac 12}a_k\prod_{j=1(j\ne k)}^N|a_k^2-
a_j^2|)^{1/2} \,,\\
h_1\varphi _k=
&
-a_k^2\varphi _k, \quad k=1,\ldots ,N
\end{array}
\]
where $W^{(k)}(u_1,\ldots ,u_N)$ is the Wronskian of the
functions
$ u_1,\ldots ,u_N $ except for the function $u_k$
and the factor $N_k$ is introduced to ensure the normalization
of the functions $\varphi _k$,
$\langle \varphi _k|\varphi _j\rangle =\delta _{kj}$,
$k,j=1,\ldots N$.
The continuous spectrum corresponds to the semiaxis $E>0$.
Continuous spectrum eigenfunctions, $\varphi _p=\varphi _p(x,t)
$ of the
Hamiltonian $h_1$ may be obtained with the aid of the operator
$L$: $\varphi _p= N_p^{-1}L\psi _p$ where the factor
$N_p^{-1}$ such that
\[N_p^2=(p^2+a_1^2)\ldots (p^2+a_N^2)\]
is introduced to ensure
the normalization of the functions $\varphi _p$:
$\langle \varphi _p|\varphi _q\rangle =\delta (p-q)$,
$h_1\varphi _p=p^2\varphi _p$.
The set $\left\{\varphi _j,\ j=1,\ldots ,N;\ \varphi _p,\ p\in
\Bbb R\right\}$ is complete in $H$.
Since the operator $L$ is linear, the relation
$L\psi _p=N_p\varphi _p$ defines the action of this operator
on every $\psi $ of the form
\begin{equation}
\psi (x,t)=\int C(p)\psi _p(x,t)dp
\label{cp}
\end{equation}
where $C(p)$ is a finite continuous function over $\Bbb R$. The
set of functions of the form (\ref{cp}) is a linear space that
I shall denote by ${\cal L}_{0p}$ and it is dense (in the sense
of generalized functions) in $H$. (More precisely it is dense
in $H_-$ since these are functionals.) Hence, the action of the
operator $L$ is defined for every element from ${\cal L}_{0p}$.
The image of the space ${\cal L}_{0p}$, that I shall denote by
${\cal L}_{1p}$ consists of the functions
\[
\varphi (x,t)=\int C(p)N_p\varphi _p(x,t)dp\, .
\]
The Darboux transformation operator $L$ together with its
Laplace
adjoint $L^+$ has remarkable factorization properties
\cite{BSTMF,AIS}
\begin{equation}
g_0=L^+L=(h_0+a_1^2)\ldots (h_0+a_N^2)\, ,
\label{fac0}
\end{equation}
\vspace{-3ex}
\begin{equation}
g_1=LL^+=(h_1+a_1^2)\ldots (h_1+a_N^2)\, .
\label{fac}
\end{equation}
The functions $\psi _p$ are eigenfunctions of $g_0$,
$g_0\psi _p=N_p^2\psi _p$. This imply that the functions
$\varphi _p$ are eigenfunctions of the operator $g_1$,
$g_1\varphi _p=N_p^2\varphi _p$.
The discrete spectrum eigefunctions of the operator $h_1$,
$\varphi _k$, $k=1,\ldots ,N$ belong to the kernel of the
operator $g_1$, $g_1\varphi _k=0$, $k=1,\ldots ,N$. This means
that the operator $g_1$ is nonnegative in $H$. Therefor,
consider the orthogonal decomposition of the space $H$:
$H=H_0\oplus H_1$ where $H_0$ is an $N$-dimensional space with
the basis $\varphi _k$, $k=1,\ldots ,N$. The functions $\varphi
_p$, $p\in \Bbb R$ form a basis (in the sense of generalized
functions) in $H_1$. In what follows I shall not consider the
space $H_0$ and restrict the consideration only by the
space $H_1$. The operators $h_1$ and $g_1$ being restricted to
this space have only a continuous spectrum and the operator
$g_1$ is strictly positive. I conserve the same notations for
these operators as operators acting in $H_1$
Taking into account the spectral decomposition of these
operators
\[h_1=\int dpp|\varphi _p\rangle \langle \varphi _p|\, ,\]
\vspace{-4ex}
\[g_1=\int dpN_p^2|\varphi _p\rangle \langle \varphi _p| \]
we can specify their domain of definitions. For the operator
$h_1$ it consists of all $\varphi \in H_1$ such that the
integral
\[ \|h_1\varphi \|^2=
\int dpp^2|\langle \varphi|\varphi _p\rangle |^2 \]
converges and for the operator $g_1$ we should demand the
convergence of the integral
\[ \|g_1\varphi \|^2=
\int dpN_p^4|\langle \varphi|\varphi _p\rangle |^2\, .\]
It is clear that the operator $g_1$ is defined on ${\cal
L}_{1p}$ and maps this space into itself. Using this fact and
the factorization property (\ref{fac}) we can define the action
of the operator $L^+$ onto
the functions $\varphi _p$,
$L^+\varphi _p=N_p^{-1}L^+L\psi _p=N_p\psi _p$, and extend this
operator by linearity on the whole space ${\cal L}_{1p}$.
It is not difficult to see that the following equality
\[\langle L\psi _p|\varphi _q\rangle =
\langle \psi _p|L^+\varphi _q\rangle \]
holds for all $\psi _p\in {\cal L}_{0p}$ and
$\varphi _q\in {\cal L}_{1p}$. Nevertheless, this fact does not
mean that $L^+$ is an operator conjugate with respect to the
scalar product to $L$ which domain of definition is ${\cal
L}_{0p}$. To find such an operator we have to specify correctly
its domain of definition.
I shall not look for this domain. Instead I shall give a closed
extension $\bar L$ of the operator $L$ and then find its
conjugate
$\bar L^+$.
Once we know the bases $\psi _p$ and $\varphi _p$ in $H$ and
$H_1$ respectively we can consider isometric operators
\[U=\int dp |\varphi _p\rangle \langle \psi _p| \,,\]
\vspace{-4ex}
\[U^{-1}=U^+=\int dp |\psi _p\rangle \langle \varphi _p| \, .\]
Similar operators have been introduced by L.D. Faddeev
\cite{Fadd}
and considered by L. Pursey \cite{Pur} for purely discrete
bases.
These operators are bounded and defined for all elements from
$H$ and $H_1$ respectively.
Consider now the following operators
\begin{equation}
\bar L=\int dp N_p|\varphi _p \rangle \langle \psi _p|\, ,
\label{Lb}
\end{equation}
\vspace{-4ex}
\begin{equation}
\bar L^+=\int dp N_p|\psi _p \rangle \langle \varphi _p|\,.
\label{Lbk}
\end{equation}
It is not difficult to specify their domains of definition. For
this purpose I use the spectral decompositions of the operator
$g_0$ and its square root
\[g_0=\int dpN_p^2 |\psi _p\rangle \langle \psi _p|\, ,\]
\vspace{-3ex}
\begin{equation}
g_0^{1/2}=\int dpN_p |\psi _p\rangle \langle \psi _p|\, .
\label {g012}
\end{equation}
It follows that
\[\|\bar L\psi \|^2=\|g_0^{1/2}\psi \|^2=
\int dp N_p^2|\langle \psi |\psi _p\rangle |^2\,.\]
This means that the domain of definition of $\bar L$ coincides
with that of $g_0^{1/2}$ and consists of all $\psi \in H$ such
that the integral in the right hand side of this equation
converges. The domain of definition of $\bar L^+$ coincides
with that of the operator $g_1^{1/2}$. It is worthwhile to
mention that these domains may be described in terms of
conditions imposed on functions that are comprised in these
domains (see for example \cite{KosSarg}) since $h_0$ and
$h_1$ are
operators bounded from below and essentially self-adjoint.
It is clear from the formulae (\ref{Lb}) and (\ref{Lbk})
that the operator $\bar L^+$
is conjugate to $\bar L$ with respect to the scalar product and
their domains of definition are well specified. Moreover,
$\bar L^{++}=\bar L$. This imply \cite{Smirn,RS} that the
operator
$\bar L$ is closed. The formulae (\ref{Lb}) and (\ref{Lbk})
give quasispectral representation of the closed operators
$\bar L$ and $\bar L^+$.
It follows from the formulae (\ref{Lb}) and (\ref{Lbk})
that $\bar L\psi _p=L\psi _p=N_p\psi _p$ and
$\bar L^+\varphi _p=L^+\varphi _p=N_p\varphi _p$. This means
that $\bar L$ is a closed extension of the operator $L$ and
$\bar L^+$ is a similar extension of the operator $L^+$
when the domains ${\cal L}_{0p}$ and ${\cal L}_{1p}$ are
taken as their initial domains of definitions.
From the spectral decomposition of the operators $g_0^{1/2}$
(\ref{g012}) and $g_1^{1/2}$,
\[g_1^{1/2}=\int dp N_p|\varphi _p\rangle
\langle \varphi _p| \,,\]
we obtain the following representations for $\bar L$ and $\bar
L^+$:
\[\bar L=Ug_0^{1/2}=g_1^{1/2}U \, ,\quad
\bar L^+=g_0^{1/2}U^+=U^+g_1^{1/2}\, .\]
Such representations are known as {\it polar decompositions} or
{\it canonical representations} of
closed operators (see for example Refs. \cite{RS,DS}).
The action of the operator $U$ on the basis $\psi _n$ gives an
orthonormal basis in $H_1$:
$\zeta _n=U\psi _n$,
$\langle \zeta _n|\zeta_k\rangle =\delta _{nk}$. The functions
$\varphi _n=g_1^{1/2}\zeta _n=\bar L\psi _n=L\psi _n$, hence,
form a basis in $H_1$ equivalent to an orthonormal (so called
Riesz basis, see for example Ref. \cite{GK}). The operator $U$
is
nonlocal and rather complicated. Therefor there is no way in
which simple explicit expressions can be derived for the
functions $\zeta _N$. The functions $\varphi _n(x,t) =L\psi
_n(x,t)$ are
much simpler but they are not orthogonal to each other:
$\langle \varphi _n|\varphi _k\rangle =S_{nk}$.
I shall denote by $S$ the infinite matrix with the entries
$S_{nk}$. The elements of this matrix can easily be expressed
in terms of the elements of another matrix $S^0(a)$ with the
entries $S_{nk}^0(a)=\langle \psi _n|h_0+a^2|\psi _k\rangle $:
\[S_{nk}=\left[ S^0(a_1)S^0(a_2)\ldots S^0(a_N) \right]_{nk}\]
where the use of the factorization property (\ref{fac0}) has
been made. Taking into account that $h_0$ is expressed in terms
of the ladder operators $a$ and $a^+$ for the basis functions
$\psi _n$,
$h_0={\textstyle \frac 14} (a+a^+)^2$, we derive the nonzero
elements of
the matrix $S^0(a)$: $S_{nn}^0(a)=n/2+1/4+a^2$,
$S_{nn+2}^0(a)={\textstyle \frac 14}\sqrt {(n+1)(n+2)}$. All
the other
matrix elements are zero. We see, hence, that the number of
nonzero elements in each row and column of the matrix $S$ is
finite.
Consider now bounded operators
\[M=\int dp N_p^{-1}|\varphi _p\rangle \langle \psi _p| \,,\]
\vspace{-3ex}
\[M^+=\int dp N_p^{-1}|\psi _p \rangle \langle \varphi _p|\]
defined in $H$ and $H_1$ respectively. It is not difficult to
see that $M\bar L^+=\Bbb I$ is the unit operator in $H_1$ and
$M^+\bar L=\Bbb I$ is the unit operator in $H$. Using the
spectral resolutions of the operators $g_0^{-1/2}$ and $g_1^{-
1/2}$
\[g_0^{-1/2}=
\int dpN_p^{-1}|\psi _p\rangle \langle \psi _p| \,,\]
\vspace{-3ex}
\[g_1^{-1/2}=
\int dpN_p^{-1}|\varphi _p\rangle \langle \varphi _p|\]
we derive the polar decompositions of the operators $M$ and
$M^+$:
\[M=Ug_0^{-1/2}=g_1^{-1/2}U\,, \]
\vspace{-3ex}
\[M^+=g_0^{-1/2}U^+=U^+g_1^{-1/2}\,.\]
It is easily seen that these operators factorise the operators
inverse to $g_0$ and $g_1$: $M^+M=g_0^{-1}$, $MM^+=g_1^{-1}$.
The functions $\eta _n=g_1^{-1/2}\zeta _n=M\psi _n$ form
another basis in $H_1$ equivalent to an orthonormal. This basis
is
biorthogonal to $\varphi _n$,
$\langle \varphi _n|\eta _k\rangle =\delta _{nk}$.
It follows the representation for the elements $S_{nk}^{-1}$
$$
\begin{array}{rl}
S_{nk}^{-1}= &
\langle \eta _n|\eta _k \rangle =
\langle \psi _n|g_0^{-1}|\psi _k \rangle \\
= &
{\displaystyle \int }dpN_p^{-2}\langle \psi _n|\psi _p\rangle
\langle \psi _p|\psi _k\rangle
\end{array}
$$
As a final remark of this section I would like to notice the
following.
The space $H_1$ can be obtained as a closure of the lineal
${\cal L}_1$ of all finite linear combinations of the functions
$\varphi _n=L\psi _n$ with respect to the norm generated by the
scalar product which is a restriction of the given scalar
product in $H$ to the lineal ${\cal L}_1$. The set of
functions of the form $\varphi =\bar L\psi $ when $\psi $ run
through the whole domain of definition of the operator $\bar L$
(i.e. the domain $D_{\sqrt {g_0}}$ of definition of the
operator $\sqrt {g_0}$) can not give the whole space $H_1$.
Nevertheless, if we define a new scalar product in
${\cal L}_1$,
$\langle \varphi _a|\varphi _b\rangle _1\equiv
\langle L\psi _a|L\psi _b\rangle =
\langle \psi _a|g_0|\psi _b\rangle $,
$\psi _{a,b}\in {\cal L}_0$, $\varphi _{a,b}\in {\cal L}_1$
then the closure of ${\cal L}_1$ with respect to the norm
generated by this scalar product will coincide with the set
$\varphi =\bar L\psi $, $\psi \in D_{\sqrt {g_0}}$. This space
is embedded in $H_1$.
\section{COHERENT STATES OF SOLITON POTENTIALS}
The operator $g_0$ is a symmetry operator for the
Schr\"o\-din\-ger equation.
Therefor it commutes with the Schr\"o\-din\-ger operator
$i\partial _t-h_0$
when applied to the solutions of the
Schr\"o\-din\-ger equation. It follows that
the operator $U=\bar Lg_0^{-1/2}$ is an intertwining operator
for the Schr\"o\-din\-ger operators $U(i\partial _t-
h_0)=(i\partial _t-h_1)U$
and therefor it is a transformation operator. Hence, being
applied to a solution of the initial
Schr\"o\-din\-ger equation (in our case the
free particle Schr\"o\-din\-ger equation) it gives a solution
of the transformed
equation (in our case the Schr\"o\-din\-ger equation with
the $N$-soliton
potential). The functions $\zeta _n=U\psi _n $ and $\zeta
_z=U\psi _z$ are then solutions of the
Schr\"o\-din\-ger equation with soliton
potential. The Fourier decomposition of the function $\zeta _z$
in terms of the basis $\left\{\zeta _n\right\}$ has the same
form as that of the function $\psi _z$ in terms of
$\left\{\psi _n\right\}$
\[\zeta _z=\Phi \sum\nolimits _n a_n\zeta _n\, .\]
The vectors $\zeta _z$, $z\in \Bbb C$ satisfy all the
conditions formulated for CS in the Introduction because of the
isometric nature of the operator $U$. The resolution of the
identity operator (\ref{Id}) in the space $H_1$ in terms of the
projectors on $\zeta _z$ takes place with the same measure
$d\mu =dxdy/\pi $, $z=x+iy$. One of the deficiencies of these
coherent states is that no a simple explicit expression for
such vectors exists. This deficiency may be cured by acting to
them by a symmetry operator for the
Schr\"o\-din\-ger equation with soliton potential.
Consider the vectors
\[\varphi _z =g_1^{1/2}\zeta _z=\bar L\psi _z=
\Phi \sum \nolimits _na_n\varphi _n\, .\]
It is not difficult to see that the value
$\langle \psi _z|g_0|\psi _z\rangle $ is finite. This means
that $\psi _z $ belong to the domain of definition of the
operator $\bar L$ and the above equality has a sense. Moreover,
these functions are sufficiently smooth and we can apply to
them directly the differential operator $L$. Thus we obtain a
coordinate representation of $\varphi _z$. For instance, in the
case of the one-soliton potential this representation reads
$$
\begin{array}{rl}
\varphi _z(x,t)= &
-{\textstyle \frac 12} (2\pi )^{-1/4}(1+it)^{-3/2} \\
\times &
[x+2iz+2a(1+it)\tanh (ax)] \\
\times &
\exp \left[ -{\displaystyle \frac {(x+2iz)^2}{4+4it}}-
{\textstyle \frac 14} (z+\bar z)^2
\right].
\end{array}
$$
We see thus that these functions are much simpler then $\zeta
_z$ and may be analyzed without difficulties. For example it is
easily seen that \cite{NF}
$|\varphi _z(x,t) |^2=|\varphi _z(-x,-t)|^2$. This property
reflects a
transparent nature of the one-soliton potential.
Another system of states may be obtained with the help of the
transformation operator $M$. Consider the vectors
\[
\eta _z=g_1^{-1/2}\zeta _z=M\psi _z=\Phi \sum\nolimits _na_n
\eta _n\, .
\]
The operator $M$ being inverse to $L$ has an integral nature.
For the case of the one-soliton potential the integration may
be carried out analytically \cite{NF}. This yields
$$
\begin{array}{rl}
\eta _z(x,t)= &
{\textstyle \frac {-i}{4}}\sqrt \pi (2\pi )^{-1/4}{\rm
sech}(ax)\\
\times &
\exp \! \left[ -{\textstyle \frac 14}(z+\bar z)^2+a^2(1+it)
\right] \\
\times &
\left[ \exp (2iaz){\rm erfc}\!
\left( a\sqrt {1+it}+{\displaystyle \frac {x/2+iz}{\sqrt
{1+it}}} \right)
\right. \\
- &
\left. \exp (-2iaz){\rm erfc}\!
\left( a\sqrt {1+it}-{\displaystyle \frac {x/2+iz}{\sqrt
{1+it}}} \right)
\right]\! .
\end{array}
$$
Where the parameter $b$ is taken to be zero.
It is worthwhile to mention that all the states
$\psi _z(x,t)$, $\varphi _z(x,t)$, $\eta _z(x,t)$, and $\zeta
_z(x,t)$
can not represent nonspreading in time wave packets.
Nevertheless, we can interpret them as coherent states since
they satisfy all the properties of such states enumerated in
the Introduction. I shall show now that for the vectors
$\varphi _z$
and $\eta _z$ there exist measures
$\mu _\varphi =\mu _\varphi (z,\bar z)$ and
$\mu _\eta =\mu _\eta (z,\bar z)$ that realize the resolution
of the identity operator in $H_1$ in terms of the projectors on
these vectors.
First consider another continuous basis in $H_1$:
$\eta _p=N_pM\psi _p$,
$\langle \eta _p|\eta _q\rangle =\delta (p-q)$,
$p,q\in \Bbb R$. Since $\left\{\varphi _p\right\}$ and
$\left\{\eta _p\right\}$ are bases in $H_1$, the resolutions of
the identity operator of the type (\ref{Id}) in terms of the
vectors $\eta _z$ and $\varphi _z $ are equivalent to the
equations
$$
\begin{array}{c}
{\displaystyle\int }d\mu _\eta (z,\bar z)\langle \eta _p|\eta
_z\rangle
\langle \eta _z|\eta _q\rangle =\delta (p-q)\, ,\\
{\displaystyle \int }d\mu _\varphi (z,\bar z)\langle \varphi
_p|\varphi _z\rangle
\langle \varphi _z|\varphi _q\rangle =\delta (p-q)\, .
\end{array}
$$
Taking into account that the functions $\psi _p$ are the
eigenfunctions of $g_0$ and $g_0^{-1}$, $g_0\psi _p=N_p^2\psi
_p$, $g_0^{-1}\psi _p=N_p^{-2}\psi _p$ we arrive at equations
for the measures $\mu _\eta $ and $\mu _\varphi $
\hspace{-6em}\begin{equation}
(N_pN_q)^{-1}\int
d\mu _\eta \langle \psi _p|\psi _z\rangle
\langle \psi _z|\psi _q\rangle =\delta (p-q)\, ,
\label{mueta}
\end{equation}
\vspace{-3ex}
\hspace{-6em}\begin{equation}
N_pN_q \int
d\mu _\varphi \langle \psi _p|\psi _z\rangle
\langle \psi _z|\psi _q\rangle =\delta (p-q)\, .
\label{mufi}
\end{equation}
Note that the integrals involved in these equations are time-%
independent and hence can by calculated at $t=0$. Therefor in
what follows I let $t=0$ and look for the measures independent
on time.
The momentum representation of the CS $\psi _z$ is well-known
\[
\begin{array}{rl}
\langle \psi _p|\psi _z\rangle =
&
(2/\pi )^{1/4}\Phi \psi _p(z)\,, \\
\psi _p(z)=
&
\exp (-p^2+2zp-z^2/2)\,,\ z=x+iy\,.
\end{array}
\]
Let us look for the measure $\mu _\eta $ in the form
$d\mu _\eta =\omega _\eta (x)dxdy$, $z=x+iy$. After performing
the integration with respect to $y$ in the Eq.
(\ref{mueta}) we arrive at an equation for $\omega _\eta (x)$
\[
(2\pi )^{1/2}\int dx \omega _\eta(x)F_p(x)=N_p^2\exp (2p^2)\,,
\]
\vspace{-3ex}
\[
F_p(x)=\exp (4px-2x^2)\,.
\]
The function $N_p^2$ is a polynomial in $p$ which is known. We
conclude then that $\omega _\eta (x)$ is a polynomial in $x$
whose coefficients are uniquely defined by the coefficients of
the polynomial $N_p^2$. For instance, for the one-soliton
potential we have
\[
\omega _\eta (x)=(x^2+a^2-1/4)/\pi\,.
\]
This proves that the states $\eta _z$ may be interpreted as CS.
We note that the states $\eta _z$ are defined with the help of
the bounded operator $g_0^{-1/2}$. This is the reason for which
the measure $\mu _\eta $ is expressed in terms of ordinary (non
generalized) functions. An other case takes place for the
states $\varphi _z $ which are defined by the semibounded
operator
$g_1^{1/2}$. I shall show now that the measure $\mu _\varphi $
is expressed in terms of generalized functions.
Let us look for the measure $\mu _\varphi $ in the form
$d\mu _\varphi =dyd\omega _\varphi (x)$. The integration in the
equation (\ref{mufi}) with respect to $y$ leads us to an
equation for the measure $d\omega _\varphi (x)$
\begin{equation}\label{Fp}
(2\pi )^{1/2}\int d\omega _\varphi (x)F_p(x)= N_p^{-2}
\exp (2p^2)\,.
\end{equation}
First we note that
$|F_p(x+iy)|\le \exp (-dx^2+by^2)$ where $2\le d\le b$.
This means that $F_p(x)$ belongs to a subspace of the space
$S_{1/2}^{1/2}$ of entire functions $F$ such that
$|F(x+iy)|\le \exp (-dx^2+by^2)$, $0\le d\le b$ \cite{GV}.
We look for $\omega _\varphi $ as a functional (i.e. a
generalized function) over $S_{1/2}^{1/2}$. (We will see that
really this is a functional over a subspace
${\stackrel{\circ}{{S}}}{}_{1/2}^{1/2}\subset
S_{1/2}^{1/2}$.)
As it is known \cite{GV} positive definite functionals (we look
for just such a functional) over $S_{1/2}^{1/2}$ are specified
by their Fourier transforms. Let $\tilde \omega _\varphi $ be
the Fourier transform of the measure $\omega _\varphi (x)$.
This means that an integration of a function
$F(x)\in S_{1/2}^{1/2}$ with respect to the measure
$\omega _\varphi (x)$ should be replaced by the integration
of the Fourier transform $\tilde F(t)$ of this function with
respect to the measure $\tilde \omega _\varphi $. In
particularly
\begin{equation}\label{Fpx}
\int d\omega _\varphi (x)F_p(x)=
\int d\tilde \omega _\varphi (t)\tilde F_p(t)
\end{equation}
where
$\tilde F_p(t)$ is the
Fourier image of the function $F_p(x)$ which in our case can
easily be found
\[\tilde F_p(t)=\sqrt {\pi /2}\exp (2p^2+ipt-t^2/8)\,.\]
As a result the Eq. (\ref{Fp}) yields the equation for
$\tilde \omega _\varphi (x)$
\[
\pi \int d\tilde \omega _\varphi (t)\exp (-t^2/8+ipt)=
N_p^{-2}\,.
\]
It is an easy exercise to see that $\tilde \omega _\varphi
(t)$
may be expressed in terms of elementary functions. For this
purpose we look for $\tilde \omega _\varphi (t)$ in the form
$d\tilde \omega _\varphi (t)=\rho _\varphi (t)dt$ and use the
following representation for the function $N_p^{-2}$:
\begin{equation}\label{Np}
\begin{array}{rl}
N_p^{-2}=
&
{\displaystyle \sum }_{k=1}^N {\displaystyle \frac {A_k}{\tau
+a
_k^2}}\,, \quad \tau=p^2\,,\\
A_k=
&
\left[(dN_p^2/d\tau )_{\tau =-a _k^2}\right]^{-1}.
\end{array}
\end{equation}
After some algebra we obtain a formula for $\rho _\varphi (t)$
\hspace{-6em}\begin{equation}\label{rofi}
\rho _\varphi (t)=
(2\pi )^{-1}\sum\nolimits _{k=1}^{N}\frac {A_k}{a _k}
\exp (t^2/8-a _k|t|)\,.
\end{equation}
Note that for the function $\rho _\varphi (t)$ of the form
(\ref{rofi}) there exist in $S_{1/2}^{1/2}$ such functions
$F(p)$ that
the integral in the right hand side of the Eq. (\ref{Fpx})
diverges. The convergence condition for this integral imposes a
restriction on the decrease of the integrand function $F(x)$
in the left hand side of the Eq. (\ref{Fpx}) as
$|x|\to \infty $. This function should satisfy an inequality
$|F(x)|\ge \exp (-2x^2-Ax)$ where $A$ is a nonnegative constant
own to every function $F(x)\in S_{1/2}^{1/2}$. I denote the set
of functions satisfying this condition by
${\stackrel{\circ }{{S}}}{}_{1/2}^{1/2}(\subset S_{1/2}^{1/2})$
which obviously is a linear space.
Thus, we have found the measure $\mu _\varphi $ in terms of the
generalized function
$\omega _\varphi (x)$ over the space
${\stackrel{\circ }{{S}}}{}_{1/2}^{1/2}$,
$d\mu_\varphi =dyd\omega _\varphi (x)$, $z=x+iy$ which is
defined by its Fourier transform $\tilde \omega _\varphi $.
The integrals with respect to this measure should be calculated
as follows
\[
\int d\mu _\varphi
\langle \varphi _a|\varphi _z\rangle
\langle \varphi _z|\varphi _b\rangle \equiv
\int dt \tilde \rho _\varphi (t)\tilde F_{ab}(t)
\]
where $ \tilde F_{ab}(t)$ is the Fourier transform of the
function
\[
F_{ab}(x)=\int dy
\langle \varphi _a|\varphi _z\rangle
\langle \varphi _z|\varphi _b\rangle \,,\
z=x+iy\,.\]
Finally I give comments on the calculation of the norms of
the functions $\eta _z$ and $\varphi _z$. The square of the
norm of $\eta _z$ may be calculated with the aid of the formula
(\ref{Np}) for the function $N_p^{-2}$ and the factorization
property of the operator $g_0^{-1}$ in terms of the operators
$M$ and $M^+$
\[
\langle \eta _z|\eta _z\rangle =
\langle \psi _z|g_0^{-1}|\psi _z\rangle =
\int dp N_p^{-2}|\langle \psi _z|\psi _p \rangle |^2\,.
\]
After some algebra we obtain
$$
\begin{array}{rl}
\langle \eta _z|\eta _z\rangle =
&
\sum \nolimits _{k=1}^NA_kF_k\,, \quad z=x+iy\,,\\
F_k=
&
{\displaystyle \frac {\sqrt {2\pi }}{a_k}}\exp [2(a_k ^2-x^2)]
\vphantom {\left({\displaystyle \frac {\sqrt {2\pi
}}{a_k}}\right)}\\
\times
&
{\rm Re}\left[\exp (4ia_kx)
{\rm erfc}(a_k \sqrt 2+i\sqrt 2x)\right]\,.
\end{array}
$$
Similarly, the square of the norm of the function $\varphi _z$
coincides with the expectation value of the operator $g_0$ in
the state $\psi _z$. For instance, for the one-soliton
potential we obtain
$\langle \varphi _z|\varphi _z\rangle =
\langle \psi _z|g_0|\psi _z\rangle =
1/4+a^2+x^2$, $z=x+iy$.
\section{CONCLUSION}
A classical particle proves no reflection in the scattering
process on a potential well. For a quantum particle in general
this is not the case. Nevertheless, there exists a wide class
of potentials called transparent potentials for which the
scattering process of the quantum particle comes in some sense
about in a similar way that those of the classical particle
i.e. without reflection. In my opinion this mysterious
phenomena up to now has no any perspicuous explanation. From a
practical point of view the answer to this question is rather
important. If at quantum level we would be able to force a
signal to propagate without reflection we could decrease the
output of the emitted signal. All transparent potentials known
at present have a remarkable property. They are related with
zero potential (free particle) by Darboux transformations. Up
to recent times it was believed that such potentials have a
finite number of discrete spectrum levels. Nevertheless a
method based on an infinite chain of Darboux transformations
with the help of which one can create transparent potentials
with infinite number of discrete spectrum levels has been
proposed recently \cite{Shab}. To understand better the nature
of transparent potentials we should investigate them in all
details.
As it is well known the quantum theory gives a more detailed
description of the nature then the classical one. Therefor
different quantum systems may correspond to the same
classical system.
Furthermore, the quantization
procedure is not unique (canonical quantization,
Berezin quantization, geometric quantization, etc.).
In this respect the following question is
of interest. What are common points between two classical
systems a quantization of which gives the quantum systems that
are related to each other by a Darboux transformation operator?
In particularly, what are common points between the classical
free particle and the particle that moves in a potential
quantization of which gives a transparent potential? The CS
approach make it possible to formulate clear steps in the
direction of obtaining an answer to this question. It permits
one to construct a classical mechanics counterpart of a given
quantum system and analyze properties of such a system. This
approach has been realized recently for the potential of the
form $x^2+gx^{-2}$ \cite{SamJMP}. It was established that at
classical level the Darboux transformation consists in a
distortion of a phase space of the classical system. Moreover,
this distortion is consistent with the transformation of the
Hamilton function in such a way that the equations of motion
remain unchanged.
Up to now no any approach for analysis of CS of transparent
potentials has been proposed. In this paper I show that the
Darboux transformation operator approach is suitable for this
purpose. A next step in this direction would be an analysis of
the classical counterpart of the quantum system that moves in a
transparent potential.
\vspace{4ex}
\begin{center}
{\bf ACKNOWLEDGMENTS}
\end{center}
It is a pleasure to thank Dr. V.P. ~Spiridonov for many helpful
discussions. This work was supported in part by the Russian
Fund for Fundamental Research and the Russian Ministry of
Education.
|
\section{Introduction}
The understanding of phase transitions on non-crystalline structures has
been recently improved by exact results connecting general geometrical features
of networks to the existence of spontaneous symmetry breaking
\cite{panza1,panza2,mwg2,fss}.
The situation is more complex when dealing with the critical behaviour of model systems.
For continuos symmetry models the singularity of the free energy, which determines the
critical behavior, appears to be related to the infrared spectrum of the
Laplacian operator on the network \cite{sfer,on}, while this is not the case for
discrete symmetries (e.g. the Ising model).
There, all known results suggest that the link between critical behaviour and geometry
should involve some other topological features \cite{panza1,panza2}.
An interesting result concerns the Sierpinski Gasket,
a typical and widely studied fractal, where the
Ising model is exactly solved. On this structure, although continuous symmetry models
exhibit a power law behaviour for $T\to 0$, the Ising model has an exponential low temperature behaviour which coincides with that found on the linear
chain \cite{panza1,panza2}.
To analyze the critical regime from an {\it ab initio} point of view, an interesting
picture is provided by the study of the singularities of thermodynamic
potentials.
In 1952 Lee and Yang in two famous papers \cite{leeyang1,leeyang2} first
proposed their fundamental approach to phase transitions, consisting in studying
the zeros of the partition function of a statistical system,
considered as a function of a complex
parameter. The partition function on a finite volume is a
polynomial in complex activity or fugacity, so that the complete
knowledge of the zeros distribution is equivalent to the knowledge of
the partition function itself and all thermodynamic quantities can be
obtained from it. On a finite volume there are no real zeros, the coefficient
being all real and positive. However in the thermodynamic limit the zeros
can pinch the real axis (the region of physical interest) producing a
singularity in the free energy (or grand-canonical potential)\cite{ruelle}.
The pinching points are phase transitions points on the
parameter axis and the zeros distribution in their neighbourhood can be
connected with the critical properties of the system \cite{ipz}.
Unfortunately, a complete knowledge of the zeros is very
difficult to obtain except for a few exactly solvable cases.
General theorems hold for the zeros distribution in the
complex magnetic field plane in a class of ferromagnetic
lattice systems, including the Ising model \cite{leeyang2,griff}.
On the other hand, very little
is known rigorously about the behaviour of the zeros of the partition function
in the complex temperature plane, the so called Fisher zeros \cite{fisher}.
In general, it is not clear if Fisher zeros arrange on smooth curves
even if this is the case in some exactly solvable models.
For the Ising model on regular two dimensional lattices \cite{fisher,shrock},
the Fisher zeros arrange on curves that cross the positive real axis at the
transition point. In the one dimensional case only two zeros
(with infinite multiplicity) are found and these have a nonzero imaginary
part, so that there is no singular point for the free energy.
For statistical models defined on non periodic discrete structures,
Fisher zeros show some peculiar features, making the analysis
of their density and location extremely subtle.
In particular, on some hierarchical lattices (i. e. q-potts model
on diamond hierarchical lattices \cite{derrida}) the zeros
have been show to form a fractal set (Julia sets). In this case, while
the general approach for identifying the singularity points and the
critical behaviour still holds,
the widely used arguments concerning scaling of singularities
and zeros density with the volume must be handled carefully,
as will be shown in the following.
In this paper we will study the Ising model on the Sierpinski gasket,
obtaining a recursive relation for the
partition function, from which the zeros of the $n$-th stage gasket can be
obtained from
those of the $(n-1)$th. The distribution we obtain is fractal and
pinches the real axis at $T=0$, so that a singular point with a power law
critical behaviour could be expected. However, since the zeros
density is found out to vanish exponentially in the neighbourhood of
$T=0$, these zeros don't produce any singularity
of the free energy: although the zeros pinch the real axis the `critical
behaviour' is the same as the one-dimensional case.
\section{Ising model on the Sierpinski gasket}
The Sierpinski gasket is a fractal graph which can be built recursively with
the following procedure: the initial stage (${\cal G}_0$) is a triangle
(3 sites with 3 edges) and the $n$th stage (${\cal G}_n$) is obtained
joining 3 ${\cal G}_{n-1}$
at their external corners, to form a bigger triangle (fig. \ref{gask}).
\begin{figure}[t]
\begin{center}
\begin{picture}(290,140)(-20,-20)
\put(0,0){\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\put(10,-20){${\cal G}_0$}
\put(50,0){\begin{picture}(60,60)
\multiput(0,0)(30,0){2}{\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture} }
\put(15,30){\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\end{picture}}
\put(75,-20){${\cal G}_1$}
\multiput(130,0)(60,0){2}{\begin{picture}(60,60)
\multiput(0,0)(30,0){2}{\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\put(15,30){\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\end{picture}}
\put(160,60){\begin{picture}(60,60)
\multiput(0,0)(30,0){2}{\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\put(15,30){\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\end{picture}}
\put(185,-20){${\cal G}_2$}
\end{picture}
\end{center}
\caption{First iterations of gasket's construction}
\label{gask}
\end{figure}
In this way ${\cal G}_n$ has $\frac{3}{2}(3^n-1)$ sites, $3^{n+1}$ edges and
its side contains $2^n$ edges.
The gasket is obtained as the limit for $n\rightarrow\infty$ of this procedure.
The Ising model on the gasket is defined associating the spin variable
$\sigma_i=\pm1$ to every site $i$ of the graph, and considering a
nearest-neighbours interaction between points joined by an edge (link).
The Hamiltonian is therefore:
\begin{equation}
E= -J\sum_{<i,j>}\sigma_i \sigma_j
\end{equation}
where the sum runs over the couples of sites joined by a link and $J$ is a
positive constant
(ferromagnetic coupling).
\section{Recursive relation for the partition function}
For ${\cal G}_0$ the partition function
\begin{equation}
Z=\sum_{\{\sigma_i \}} e^{- \beta E}
\end{equation}
can be seen as a sum of the elements of the rank 3 tensor $M_0$
\begin{equation}
Z_0=\sum_{\sigma_1,\sigma_2,\sigma_3=\pm 1} M_0^{\sigma_1\sigma_2\sigma_3}
\end{equation}
where
\begin{equation}
M_0^{\sigma_1\sigma_2\sigma_3}=\exp[-\beta E(\sigma_1,\sigma_2,\sigma_3)]
\end{equation}
$M_0^{\sigma_1\sigma_2\sigma_3}$ can take only 2 values because
there are
only 2 classes of spin configurations with different energy:
\begin{equation}
\left\{ \begin{array}{l}
M_0^{\sigma\sigma\sigma}=e^{3 \beta J}=y^3 \\
M_0^{\sigma\sigma(-\sigma)}=M_0^{\sigma(-\sigma)\sigma}
M_0^{(-\sigma)\sigma\sigma}=e^{- \beta J}=y^{-1}
\end{array}
\right.
\end{equation}
where $y= e^{\beta J}$.
In terms of $y$ the partition function is:
\begin{equation}
Z_0=2y^3+6y^{-1}
\end{equation}
\begin{figure}[t]
\begin{center}
\begin{picture}(220,100)(-20,-20)
\put(0,0){\begin{picture}(60,60)
\multiput(0,0)(60,0){2}{\circle*{6}}
\put(30,60){\circle*{6}}
\put(0,0){\line(1,0){60}}
\put(0,0){\line(1,2){30}}
\put(30,60){\line(1,-2){30}}
\end{picture}}
\put(28,70){1} \put(-12,-12){2} \put(68,-12){3}
\put(120,0){\begin{picture}(60,60)
\multiput(0,0)(30,0){2}{\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\put(15,30){\begin{picture}(30,30)
\multiput(0,0)(30,0){2}{\circle*{6}}
\put(15,30){\circle*{6}}
\put(0,0){\line(1,0){30}}
\put(0,0){\line(1,2){15}}
\put(15,30){\line(1,-2){15}}
\end{picture}}
\end{picture}}
\put(148,70){1} \put(108,-12){2} \put(188,-12){3}
\put(123,32){a} \put(173,32){b} \put(148,-12){c}
\end{picture}
\end{center}
\caption{Labeling of sites used for ${\cal G}_0$ and ${\cal G}_1$}
\label{gasklab}
\end{figure}
For ${\cal G}_1$ the partition function can be expressed in the same way
separating the sum over the states of the inner sites using the
tensor $M_1$ whose indices correspond to the spins on the external vertices
(fig. \ref{gasklab}):
\begin{equation}
M_1^{\sigma_1\sigma_2\sigma_3}=
\sum_{\sigma_a,\sigma_b,\sigma_c=\pm 1} \exp[-\beta E(\sigma_i)] =
\sum_{\sigma_a,\sigma_b,\sigma_c=\pm 1}
M_0^{\sigma_1\sigma_a\sigma_b} M_0^{\sigma_a\sigma_2\sigma_c}
M_0^{\sigma_b\sigma_c\sigma_3}
\end{equation}
Now $M_1$ has the same structure as $M_0$ since the possible values are:
\begin{equation}
\left\{ \begin{array}{l}
M_1^{\sigma\sigma\sigma}=4y^{-3}+3y+y^9 \\
M_1^{\sigma\sigma(-\sigma)}=M_1^{\sigma(-\sigma)\sigma}
M_1^{(-\sigma)\sigma\sigma}=3y^{-3}+4y+y^5
\end{array}
\right.
\end{equation}
One can obtain $M_1$ from $M_0$ simply by a transformation mapping $y^3$
in $4y^{-3}+3y+y^9$ and $y^{-1}$ in $3y^{-3}+4y+y^5$.
This is done by the substitution
\begin{equation}
y \to f(y)=\left( \frac{y^8-y^4+4}{y^4+3} \right)^{\frac{1}{4}}
\label{tr1}
\end{equation}
followed by the multiplication by
\begin{equation}
c(y)=\frac{y^4+1}{y^3} \left[ (y^4+3)^3(y^8-y^4+4)\right]^{\frac{1}{4}}
\label{tr2}
\end{equation}
The transformation also gives the new partition function:
\begin{equation} Z_1(y)=Z_0(f(y))\ c(y)
\end{equation}
Following the same argument one can obtain for the $(n+1)$th stage
of the gasket ${\cal G}_{n+1}$:
\begin{equation}
Z_{n+1}(y)=Z_n(f(y))\cdot \left[c(y)\right]^{3^n}
\end{equation}
Using this recursion relation we get:
\begin{equation}
Z_n(y)=\frac{2}{y^{3^n}} P_n(y^4)
\end{equation}
where $P_n(t)$ is a polynomial in $t$ of degree $3^n$ in which the
coefficient of $t^{3^n}$ is 1;
for $n=0$ one has $P_0(t)=t+3$
while the general case $n>0$ can be proven by induction.
\section{Zeros of the partition function}
Introducing the variable $x = y^4$ the transformation (\ref{tr1}), (\ref{tr2})
is given by:
\begin{equation}
\left\{ \begin{array}{l}
x \rightarrow \tilde{f}(x)=\frac{x^2-x+4}{x+3} \\
\tilde{c}(x)=(x+1)x^{-\frac{3}{4}}
\left[(x+3)^3(x^2-x+4)\right]^{\frac{1}{4}}
\end{array}
\right.
\end{equation}
Denoting by $x_n^i$ the zeros of $P_n(x)$ the partition function reads
\begin{equation}
Z_n=\frac{2}{x^{3^n/4}} \prod_{i=1}^{3^n}(x-x_n^i)
\end{equation}
and using the recurrence one finds
\begin{equation}
2 x^{-\frac{3^{n+1}}{4}} \prod_{i=1}^{3^{n+1}}(x-x_{n+1}^i)=
2 x^{-\frac{3^{n+1}}{4}} \prod_{i=1}^{3^n} \left\{ [(x^2-x+4)-x_n^i(x+3)]
(x+1) \right\}
\end{equation}
This equation shows that for every root $x_n^i$ of $Z_n$, $Z_{n+1}$
has the root $x=-1$ and the 2 solutions of
\begin{equation}
x_n^i=\tilde{f}(x_{n+1}^i)
\end{equation}
namely the preimages of $x_n^i$ by the transformation $\tilde{f}$ .
Starting from $x_0^1=-3$
one obtains all the zeros of the partition function for the $n$-th stage
gasket as shown in table~\ref{zerij}, where $h(x)$ denotes the set of
the preimages of $x$ (and $h^k(x)$ is a set of $2^k$ zeros).
\begin{table}[ht]
\[ \begin{array}{|c|l|}
\hline
n & \multicolumn{1}{c|}{\mbox{zeros}} \\ \hline \hline
0& -3 \\ \hline
1& -1\hspace{.5cm} h(-3) \\ \hline
2& -1\hspace{.5cm} h(-1)\hspace{.5cm} h(h(-3)) \\ \hline
3& -1\hspace{.5cm} h(-1)\hspace{.5cm} h(h(-1))\hspace{.5cm}
h(h(h(-3))) \\ \hline
n& -1\hspace{.5cm} h(-1)\hspace{.5cm} \ldots \hspace{.5cm}
h^{n-1}(-1)\hspace{.5cm} h^n(-3) \\ \hline
\end{array} \]
\caption{Zeros of partition function in y}
\label{zerij}
\end{table}
From this table one can see that the preimages of $-1$ by the $j$th iterate
of $\tilde{f}$ appear at the $(j-1)$th stage and are zeros of all the
following stages, while the preimages of $-3$ are `temporary' zeros.
It is also possible to find the multiplicity of these zeros: in fact since
every
root generates the root $-1$ at the next stage, this value appears $3^{n-1}$
times among the zeros of $n$-th stage, the roots $h(-1)$ appear as many times
as $-1$ in the previous stage (their multiplicity is $3^{n-2}$) and in general
the multiplicity of the roots belonging to $h^j(-1)$ is $3^{n-j-1}$.
In this way one can, in principle, calculate all the zeros of the
partition function
at any stage. In practice this is possible only for small $n$ because of
their exponential growth.
An alternative approach \cite{derrida} is to start from a root
$x_0$ (for example `$-1$' whose
preimages are `permanent') then choose at random one of its two preimages
by the transformation $\tilde{f}$ (denoted by $x_1$), then choose one of the
preimages of $x_1$ and so on; the set of points obtained in this way is a
representative of the set of all roots and has the advantage to contain zeros
relative to large $n$.
By plotting in the complex temperature plane the roots obtained with both
methods it can be seen that very few of them fall in the neighborhood of the
real axis and no information can be obtained about the critical behaviour (see
fig. \ref{rand} where the zeros are plotted in the plane of the variable
$t=e^{-\beta J}$).
\begin{figure}
\begin{center}
\mbox{\epsfig{file=fig3.eps,height=8cm,angle=90}}
\end{center}
\caption{20000 zeros obtained by the random method (in the plane of
$t=e^{-\beta J}$)}
\label{rand}
\end{figure}
A good technique to obtain more zeros near the real axis consists in changing the
choice probability of the two preimages \cite{derrida}; the two solutions of
$x=\tilde{f}(x^{\prime})$ are
\begin{equation}
x^{\prime}=h_1(x)=\frac{1 + x - {\sqrt{-15 + 14\,x + {x^2}}}}{2}
\end{equation}
\begin{equation}
x^{\prime}=h_2(x)=\frac{1 + x + {\sqrt{-15 + 14\,x + {x^2}}}}{2}
\end{equation}
and one can see that the repeated application of $h_2$ gives a sequence
of points approaching the real axis. The set of roots obtained by increasing
the probability of choosing the second preimage is not a representative
set af all roots (it doesn't show their density not even approximately) but
gives us a chance to observe their behaviour in the interesting area (see
fig. \ref{rand98}).
\begin{figure}
\begin{center}
\mbox{\epsfig{file=fig4.eps,height=8cm,angle=90}}
\end{center}
\caption{20000 zeros obtained with probability 0.98 of choosing $h_2$
(in the $t$ plane)}
\label{rand98}
\end{figure}
A plot of these roots in the plane of $w=e^{-4\beta J}$ ($T=0$ corresponds to
$w=0$) with a {\it log-log} scale (fig. \ref{loglog}) shows that the real and
imaginary part are related by a power law: the curve can intersect the real
axis in the thermodynamic limit only at the origin.
\begin{figure}
\begin{center}
\mbox{\epsfig{file=fig5.eps,height=8cm,angle=90}}
\end{center}
\caption{Log-log plot of zeros (in the variable $w=e^{-4\beta J}$)}
\label{loglog}
\end{figure}
Analytically one can verify this power behaviour by studying the
transformation of the variable $w=e^{-4\beta J}=x^{-1}$
\begin{equation}
g(w)=\left. \frac{1}{f(x)} \right|_{x=\frac{1}{w}}=\frac{w(3w+1)}{4w^2-w+1}
\label{g}
\end{equation}
Assuming that $\Im(w)=A\ \Re(w)^b$, that is
\begin{equation}
w=\xi+i A \xi^b
\label{a}
\end{equation}
and inserting (\ref{a}) in (\ref{g}) one obtains
\begin{equation}
\Im(g(\xi+i A \xi^b))= A \xi^b \left(1+8\xi+ O(\xi^{\min\{3,2b-1\}})\right)
\end{equation}
and
\begin{equation}
A\ \left(\Re(g(\xi+i A \xi^b))\right)^b=A \xi^b \left(1+4b\xi
O(\xi^{\min\{3,2b-1\}}))\right)
\end{equation}
Choosing $b=2$, one sees that the curve $w=\xi+i A \xi^2$ is `conserved'
by transformation $g$ except for higher order terms in $\xi$.
\section{Density of zeros}
We have seen that the zeros pinch the real axis only at $T=0$ and we
proceed by studying their density in the neighbourhood of this point
to establish the critical behaviour; it is important to see whether
this density (which is quite small, as we have seen) goes to zero or
remains finite.
A numerical estimate can be obtained by simply counting the zeros (with
their multiplicity).
This has been done in two ways:
\begin{itemize}
\item considering only the zeros in the neighbourhood of the real axis
and grouping them with regard to their real part (one-dimensional density);
\item dividing the complex plane in equal rectangles and counting the zeros
contained (two-dimensional density).
\end{itemize}
\begin{figure}
\begin{center}
\mbox{\epsfig{file=fig6.eps,height=8cm,angle=90}}
\end{center}
\caption{Density of zeros vs. their real part in the $t$ plane (I method)}
\label{dens1}
\end{figure}
\begin{figure}
\begin{center}
\mbox{\epsfig{file=fig7.ps,height=11cm,angle=270}}
\end{center}
\caption{Density of zeros in the $t$ plane (log scale)}
\label{dens2}
\end{figure}
\begin{figure}
\begin{center}
\mbox{\epsfig{file=fig8.ps,height=11cm,angle=270}}
\end{center}
\caption{3-dimensional view of the zeros density in the $t$ plane}
\label{dens3}
\end{figure}
In the first case one obtains plots like fig. \ref{dens1} which
shows the density of the zeros of the partition function with
$\Im(t)<.3$ for gaskets from ${\cal G}_{10}$ to ${\cal G}_{20}$;
in the second case obtains the result shown in figures \ref{dens2} and
\ref{dens3} (which refers to ${\cal G}_{18}$).
From fig. \ref{dens1} one can
see that going from one stage to the next the density does not change
appreciably except for the tail towards $0$ that grows longer but is
strongly decreasing: the density at $T=0$ appears to vanish exponentially.
This behaviour can also be verified by an analytical estimate. First we
notice that the zeros near $T=0$ are those obtained by the repeated
application of $h_2$.
Indeed for $|x|\rightarrow \infty$ we have
\begin{equation}
h_1(x) \to k
\end{equation}
while
\begin{equation} h_2(x) \to x+4-\frac{16}{x}+O(x^{-2})
\end{equation}
and, in terms of real and imaginary part,
\begin{equation} h_2(u+i v)\approx \left( u+4-\frac{16 u}{u^2+v^2}\right)+
i \left(v-\frac{16 v}{u^2+v^2}\right)
\end{equation}
Applying $h_2$ to $z=u+i~v$ for large $u$ one obtains $h_2(z) \simeq 4u+i~v$.
In this limit the density of zeros in the $x$ plane becomes
the product of two factors, one depending only on $u$ and
the other on $v$:
\begin{equation}
d(u,v) \approx d_1(u) d_2(v)
\end{equation}
where $d_2(v)$ is bounded.
The asymptotic behaviour of $d_1(u)$ for $u\to\infty$ can be obtained
by noting that, for each set $h^k(-1)$, the
zeros with real part $u$ are those obtained by a sequence ending with
the application of $h_1$ followed by $h_2^n$, where
$n=\frac{u}{4}+c$ and $c$ is a constant independent of $u$.
So this fraction of zeros is $\frac{1}{2}$ to the
power $\frac{u}{4}+c$, that is proportional to $\exp (-u/U)$ with
$U=4 \ln 2$.
Since the total density $d_1$ is a weighted sum of the partial densities that
have the same behaviour we have:
\begin{equation}
d_1(u)\propto \exp(-u/U)
\end{equation}
To find the density in the $t$ plane we must now divide $d$ by the Jacobian
of the transformation:
\begin{equation}
t=x^{-\frac{1}{4}}
\end{equation}
This Jacobian turns out to be
\begin{equation}
\frac{1}{16} (u^2+v^2)^{-\frac{5}{4}}
\end{equation}
and finally
\begin{equation}
\tilde{d}(t_r,t_i)\propto\left. d_2(v) e^{-u/U}
(u^2+v^2)^{\frac{5}{4}} \right|_{t_r,t_i}
\end{equation}
where $t_r$ e $t_i$ are the real and imaginary part of $t$.
For $t\rightarrow0$ (that is $u\rightarrow +\infty$) the density
vanishes exponentially, as we could infer from numerical calculation,
and this behaviour has the same effect as a gap near the real axis.
Therefore the low temperature regime is not affected by the zeros contained in this
region and one observe a situation analogous to the one dimensional case.
This can be seen, for example, by comparing the behaviour of thermodynamical
quantities:
figure \ref{comp} shows that, even if the zeros distributions
seem to be quite
different, the behaviour of the specific heat for the Sierpinski gasket and
the linear chain is essentially the same.
\begin{figure}[p]
\begin{center}
\mbox{\epsfig{file=fig9.eps,height=12cm}}
\end{center}
\caption{Comparison between the zeros distribution in the $t$ plane and
the specific heat for the linear chain (on the left) and the gasket
(on the right)}
\label{comp}
\end{figure}
\section{Conclusions}
The anomalous behaviour of the density of zeros for the Ising model
on the Sierpinski gasket is to be deeply related to the its self-similar
geometry. This strongly suggests a careful approach to the analysis
of scaling of zeros density on fractals. In particular
a stimulating open problem is the relation of this scaling with the
geometry of a generic self-similar structure and with
known anomalous dimensions. An important step in this direction would
be the study of Fisher zeros on the more complex case of a Sierpinski
carpet, where an exact solution is still lacking but the Ising
model is expected to have a phase transition at finite temperature.
|
\section{Introduction: Distinguishability of superoperators}
The exciting recent developments in the theory of quantum information and
computation have already established an enduring legacy. The two most
far-reaching results --- that a quantum computer (apparently) can solve
problems that will forever be beyond the reach of classical computers
\cite{qc}, and that quantum information can be protected from errors if
properly encoded \cite{qec} --- have surely earned a prominent place at the
foundations of computer science.
The implications of these ideas for the future of physics are less clear, but
we expect them to be profound. In particular, we anticipate that our deepening
understanding of quantum information will lead to new strategies for pushing
back the boundaries of quantum-limited measurements. Quantum entanglement,
quantum error correction, and quantum information processing can all be
exploited to improve the information-gathering capability of physics
experiments.
In a typical quantum-limited measurement, a classical signal is conveyed over a
quantum channel \cite{mabuchi}. Nature sends us a message, such as the value of
a weak force, that can be regarded as a classical parameter appearing in the
Hamiltonian of the apparatus (or more properly, if there is noise, its master
equation). The apparatus undergoes a quantum operation $\$(a)$, and we are to
extract as much information as we can about the parameter(s) $a$ by choosing an
initial preparation of the apparatus, and a positive-operator-valued measure
(POVM) to read it out. Quantum information theory should be able to provide a
theory of the {\sl distinguishability of superoperators}, a measure of how much
information we can extract that distinguishes one superoperator from another,
given some specified resources that are available for the purpose. This
distinguishability measure would characterize the inviolable limits on
measurement precision that can be achieved with fixed resources.
Many applications of quantum information theory involve the problem of
distinguishing nonorthogonal quantum {\sl states}. For example, a density
operator $\rho_a $ is chosen at random from an ensemble ${\cal
E}=\{\rho_a,p_a\}$ (where $p_a$ is an {\it a priori} probability), and a
measurement is performed to extract information about which $\rho_a$ was
chosen. The problem of distinguishing {\sl superoperators} is rather
different, but the two problems are related. For example, let us at first
ignore noise, and also suppose that the classical force we are trying to detect
is static. Then we are trying to identify a particular time-independent
Hamiltonian $H_a$ that has been drawn from an ensemble $\{H_a,p_a\}$. We may
choose a particular initial pure state $|\psi_0\rangle$, and then allow the
state to evolve, as governed by the unknown Hamiltonian, for a time $t$; our
ensemble of possible Hamiltonians generates an ensemble of pure states
\begin{equation}
\{|\psi_a(t)\rangle = e^{-itH_a}|\psi_0\rangle, p_a\}~.
\end{equation}
Since our goal is to gain as much information as possible about the applied
Hamiltonian, we should choose the initial state $|\psi_0\rangle$ so that the
resulting final states are maximally distinguishable.
There are many variations on the problem, distinguished in part by the
resources we regard as most valuable. We might have the freedom to chose the
elapsed time as we please, or we might impose constraints on $t$. We might
have the freedom to modify the Hamiltonian by adding an additional ``driving''
term that is under our control. We might use an {\sl adaptive} strategy, where
we make repeated (possibly weak) measurements, and our choice of initial state
or driving term in later measurements takes into account the information
already collected in earlier measurements \cite{wiseman_adaptive}.
Imposing an appropriate cost function on resources is an important aspect of
the formulation of the problem, particularly in the case of the detection of a
static (DC) signal. For example, we could in principle repeat the measurement
procedure many times to continually improve the accuracy of our estimate. In
this respect, the problem of distinguishing superoperators does not have quite
so fundamental a character as the problem of distinguishing states, as in the
latter case the no-cloning principle \cite{no_clone} prevents us from making
repeated measurements on multiple copies of the unknown state. But for a
time-dependent signal that stays ``on'' for a finite duration, there will be a
well-defined notion of the optimal strategy for distinguishing one possible
signal from another, once our apparatus and its coupling to the classical
signal have been specified. Still, for the sake of simplicity, we will mostly
confine our attention here to the case of DC signals.
We don't know exactly what shape this nascent theory of the distinguishability
of superoperators should take, but we hope that further research can promote
the development of new strategies for performing high-precision measurements.
On the one hand we envision a program of research that will be relevant to real
laboratory situations. On the other hand, we seek results that are to some
degree robust and general (not tied to some particular model of decoherence, or
to a particular type of coupling between quantum probe and classical signal).
Naturally, there is some tension between these two central desiderata; rather
than focus on a specific experimental context, we lean here toward more
abstract formulations of the problem.
Our discussion is far from definitive; its goal is to invite a broader
community to consider these issues. We will mostly be content to observe that
some familiar concepts from the theory of quantum information and computation
can be translated into tools for the measurement of classical forces. Some
examples include superdense coding, fast database search, and the quantum
Fourier transform.
Naturally, the connections between quantum information theory and precision
measurement have been recognized previously by many authors. Especially
relevant is the work by Wootters \cite{wootters}, by Braunstein\cite{braun},
and by Braunstein and Caves \cite{braunstein} on state distinguishability and
parameter estimation, and by Braginsky and others \cite{braginsky} on quantum
nondemolition measurement. Though what we have to add may be relatively modest,
we hope that it may lead to further progress.
\section{Superdense coding: improved distinguishability through entanglement}
\label{sec:superdense}
Recurring themes of quantum information theory are that entanglement can be a
valuable resource, and that entangled measurements sometimes can collect more
information than unentangled measurements. It should not be surprising, then,
if the experimental physicist finds that the best strategies for detecting a
weak classical signal involve the preparation of entangled states and the
measurement of entangled observables.
Suppose, for example, that our apparatus is a single-qubit, whose
time-independent Hamiltonian (aside from an irrelevant additive constant), can
be expressed as
\begin{equation}
H_{\vec a}=\vec a\cdot \vec \sigma~;
\end{equation}
here $\vec a=(a_1,a_2,a_3)$ is an unknown three-vector, and $\sigma_{1,2,3}$
are the Pauli matrices. (We may imagine that a spin-${1\over 2}$ particle with
a magnetic moment is employed to measure a static magnetic field.) By preparing
an initial state of the qubit, allowing the qubit to evolve, and then
performing a single measurement, we can extract at best one bit of information
about the magnetic field (as Holevo's theorem \cite{holevo} ensures that the
optimal POVM in a two-dimensional Hilbert space can acquire at most one bit of
information about a quantum state).
If we have two qubits, and measure them one at a time, we can collect at best
two bits of information about the magnetic field. In principle, this could be
enough to distinguish perfectly among four possible values of the field. In
practice, for a generic choice of four Hamiltonians labeled by vectors $\vec
a^{(1,2,3,4)}$, the optimal information gain cannot be achieved by measuring
the qubits one at a time. Rather a better strategy exploits quantum
entanglement.
An improved strategy can be formulated by following the paradigm of superdense
coding \cite{wiesner}, whereby shared entanglement is exploited to enhance
classical communication between two parties. To implement superdense coding,
the sender (Alice) and the receiver (Bob) use a shared Bell state
\begin{equation}
|\phi^+\rangle= {1\over\sqrt{2}}\left(|00\rangle + |11\rangle\right)
\end{equation}
that they have prepared previously. Alice applies one of the four unitary
operators $\{I,\sigma_1,\sigma_2,\sigma_3\}$ to her member of the entangled
pair, and then sends it to Bob. Upon receipt, Bob possesses one of the four
mutually orthogonal Bell states\begin{eqnarray}
\label{bellbasis}
|\phi^+\rangle & = & {1\over\sqrt{2}}\left(|00\rangle+|11\rangle\right) = I
\otimes I |\phi^+\rangle ~,\cr
|\psi^+\rangle & = & {1\over\sqrt{2}}\left(|01\rangle+|10\rangle\right) =
\sigma_1 \otimes I |\phi^+\rangle ~,\cr
-i|\psi^-\rangle & = & {-i\over\sqrt{2}}\left(|01\rangle-|10\rangle\right) =
\sigma_2 \otimes I |\phi^+\rangle ~,\cr
|\phi^-\rangle & = & {1\over\sqrt{2}}\left(|00\rangle-|11\rangle\right) =
\sigma_3 \otimes I |\phi^+\rangle~;\end{eqnarray}
by performing an entangled Bell measurement (simultaneous measurements of the
commuting collective observables $\sigma_1\otimes\sigma_1$ and
$\sigma_3\otimes\sigma_3$), Bob can perfectly distinguish the states. Although
only one qubit passes from Alice to Bob, two classical bits of information are
transmitted and successfully decoded. In fact, this enhancement of the
transmission rate is optimal -- with shared entanglement, no more than two
classical bits can be carried by each transmitted qubit \cite{hausladen}.
The lesson of superdense coding is that entanglement can allow us to better
distinguish
operations on quantum states, and we may apply this method to the problem of
distinguishing Hamiltonians.\footnote{This idea was suggested to us by Chris
Fuchs \cite{fuchs_private}.} Let us imagine that the magnitude of the magnetic
field is known, but not its direction -- then we can choose our unit of time so
that $|\vec a|=1$. We may prepare a pair of qubits in the entangled state
$|\phi^+\rangle$, and expose only one member of the pair to the magnetic field
while the other remains well shielded. In time $t$, the state evolves to
\begin{eqnarray}
|\psi_{\hat a}(t)\rangle & \equiv &\exp\left(-itH_{\hat a}\otimes
I\right)|\phi^+\rangle \cr
& = & \left[\cos t(I\otimes I) -i \sin t(\hat a\cdot\vec\sigma\otimes I)
\right]|\phi^+\rangle\cr
& = & \cos t |\phi^+\rangle \cr
&&-i\sin t
\left[a_1|\psi^+\rangle -ia_2|\psi^-\rangle + a_3|\phi^+\rangle\right]~;
\end{eqnarray}
the inner product between the states arising from Hamiltonians $H_{\hat a}$ and
$H_{\hat b}$ becomes
\begin{equation}
\label{superip}
\langle\psi_{\hat a}(t)|\psi_{\hat b}(t)\rangle = \cos^2 t + (\hat a \cdot \hat
b) \sin^2 t~.
\end{equation}
For these states to be orthogonal, we require
\begin{equation}
\hat a \cdot \hat b = -\cot^2 t~.
\end{equation}
Since $\cot^2 t \ge 0$, the states are not orthogonal for any value of $t$
unless the two magnetic field directions $\hat a$ and $\hat b$ are separated by
at least $90^\circ$.
Now suppose that the magnetic field (of known magnitude) points in one of three
directions that are related by three-fold rotational symmetry. These
directions could form a planar trine with
$\hat a\cdot \hat b=\hat a\cdot \hat c= \hat b\cdot \hat c = -1/2$, or a
``lifted trine'' with angle $\theta$ between each pair of directions, where
$-1/2\le \cos\theta \le 0$. For any such trine of field directions, we may
evolve for a time $t$ such that
\begin{equation}
\cot^2 t = -\cos\theta ~,
\end{equation}
and perform an (entangled) orthogonal measurement to determine the field. At
the point of tetrahedral symmetry, $\cos\theta=-1/3$, we may add a fourth field
direction such that
the inner product for each pair of field directions is $-1/3$; then all four
directions can be perfectly distinguished by Bell measurement.
In this case of four field directions with tetrahedral symmetry, the two-bit
measurement outcome achieves a two-bit information gain, if the four directions
were equally likely {\it a priori}. In contrast, no adaptive strategy in which
single qubits are measured one at a time can attain a two-bit information gain.
This separation between the information gain attainable through entangled
measurement and that attainable through adaptive nonentangled measurement, for
the problem of distinguishing Hamiltonians, recalls the analogous separation
noted by Peres and Wootters \cite{peres} for the problem of distinguishing
nonorthogonal states.
\section{Grover's database search: improved distinguishability through driving}
Another instructive example is Grover's method \cite{grover} for searching an
unsorted database, which (as formulated by Farhi and Gutmann \cite{farhi}) we
may interpret as a method for improving the distinguishability of a set of
Hamiltonians by adding a controlled driving term.
Consider an $N$-dimensional Hilbert space with orthonormal basis
$\{|x\rangle\}, ~x=0,1,2,\dots,N-1$, and suppose that the Hamiltonian for this
system is known to be one of the $N$ operators
\begin{equation}
H_x=E|x\rangle\langle x| ~.
\end{equation}
We are to perform an experiment that will allow us to estimate the value of
$x$.
We could, for example, prepare the initial state ${1\over \sqrt{2}}(|y\rangle +
|y'\rangle)$, allow the system to evolve for a time $T=\pi/E$, and then perform
an orthogonal measurement in the basis $|\pm\rangle={1\over \sqrt{2}}(|y\rangle
\pm |y'\rangle)$. Then we will obtain the outcome $|-\rangle$ if and only if
one of $y,y'$ is $x$. Searching for $x$ by this method, we would have to
repeat the experiment for O($N$) distinct initial states to have any
reasonable chance of successfully inferring the value of $x$.
Our task becomes easier if we are able to modify the Hamiltonian by adding a
term that we control to drive the system. We choose the driving term to be
\begin{equation}
\label{grover_drive}
H_D=E|s\rangle\langle s|~,
\end{equation}
where $|s\rangle$ denotes the state
\begin{equation}
|s\rangle={1\over \sqrt{N}}\sum_{y=0}^{N-1}|y\rangle~.
\end{equation}
Then the full Hamiltonian is
\begin{equation}
H'_x=H_x+H_D=E(|x\rangle\langle x| + |s\rangle\langle s|)~,
\end{equation}
and we can readily verify that the vectors
\begin{equation}
|E_{\pm}\rangle\equiv |s\rangle \pm |x\rangle
\end{equation}
are (unconventionally normalized!) eigenstates of $H$ with the eigenvalues
\begin{equation}
E_{\pm}=E\left(1\pm {1\over\sqrt{N}}\right)~.
\end{equation}
We may prepare the initial state
\begin{equation}
|s\rangle= {1\over 2}(|E_+\rangle + |E_-\rangle)~;
\end{equation}
since the energy splitting is $\Delta E=2E/\sqrt{N}$, after a time
\begin{equation}
T=\pi/\Delta E= \pi\sqrt{N}/2E~,
\end{equation}
this state flops to the state
\begin{equation}
{1\over 2}(|E+\rangle - |E-\rangle)=|x\rangle~.
\end{equation}
Thus, by performing an orthogonal measurement, we can learn the value of $x$
with certainty \cite{farhi}.
The driving term we have chosen is the continuous time analog of the iteration
employed by Grover \cite{grover} for rapid searching. And as the Grover search
algorithm can be seen to be optimal, in the sense that a marked state can be
identified with high probability with the minimal number of oracle calls
\cite{bbbv}, so the driving term we have chosen is optimal in the sense that it
enables us to identify the value of the classical parameter labeling the
Hamiltonian in the minimal time, at least asymptotically for $N$ large. (In a
physics experiment, the ``oracle'' is Nature, whose secrets we are eager to
expose.) For this Grover-Farhi-Gutmann problem, we can make a definite
statement about how to optimize expenditure of a valuable resource (time) in
the identification of a system Hamiltonian.
We also note that adding a driving term can sometimes improve the efficacy of
the superdense coding method described in \S\ref{sec:superdense}. For example,
in the case of three magnetic fields of equal magnitude with threefold
symmetry, but with an
angle between fields of less than $90^\circ$, applying a driving field along
the line of
symmetry can make the resultant field directions perfectly distinguishable.
In fact, Beckman \cite{beckman} has shown that for any three field vectors
forming a triangle that is isosceles or nearly isosceles, a suitable driving
field can always by found such that the field directions can be distinguished
perfectly.
\section{Distinguishing two alternatives}
Let's consider the special case in which our apparatus is known to be governed
by one of two possible Hamiltonians $H_1$ or $H_2$. If the system is two
dimensional, we are trying to distinguish two possible values $\vec a,\vec b$
of the magnetic field with a spin-${1\over 2}$ probe. Suppose for simplicity
that the two fields have the same magnitude (normalized to unity), but
differing directions.
Assuming that we are unable to modify the Hamiltonian by adding a driving term,
the optimal strategy is to choose an initial polarization vector that bisects
the two field directions $\hat a, \hat b$. Depending on the actual value of the
field, the polarization will precess on one of two possible cones. If the angle
between $\hat a$ and $\hat b$ is $\theta\ge 90^\circ$, then the two possible
polarizations will eventually be back-to-back; an orthogonal measurement
performed at that time will distinguish $\hat a$ and $\hat b$ perfectly. But if
$\theta < 90^\circ$, the two polarizations are never back-to-back; the best
strategy is to wait until the angle between the polarizations is maximal, and
to then perform the orthogonal measurement that best distinguishes them. We
cannot perfectly distinguish the two field directions by this method.
On the other hand, if we are able to apply a known driving magnetic field in
addition to the unknown field that is to be determined, then two fields $\vec
a$ and $\vec b$ can always be perfectly distinguished. If we apply the field
$-\vec b$, then the problem is one of distinguishing the trivial Hamiltonian
from
\begin{equation}
H_{\rm diff}=(\vec a - \vec b)\cdot \vec\sigma~.
\end{equation}
We can choose an initial polarization orthogonal to $\vec a - \vec b$, and wait
just long enough for $H_{\rm diff}$ to rotate the polarization by $\pi$. Then
an orthogonal measurement perfectly distinguishes $H_{\rm diff}$ from the
trivial Hamiltonian.
Evidently, the same strategy can be applied to distinguish two Hamiltonians
$H_1$ and $H_2$ in a Hilbert space of arbitrary dimension. We drive the system
with $-H_2$; then to distinguish the trivial Hamiltonian from $H_1-H_2$, we
chose the initial state
\begin{equation}
{1\over\sqrt{2}}\left(|E_{\rm min}\rangle + |E_{\rm max}\rangle\right)~,
\end{equation}
where $E_{\rm min},E_{\rm max}$ are the minimal and maximal eigenvalues of
$H_1-H_2$. After a time $t$ with
\begin{equation}
t(E_{\rm max}-E_{\rm min})=\pi~,
\end{equation}
this state evolves to the orthogonal state ${1\over\sqrt{2}}\left(|E_{\rm
min}\rangle - |E_{\rm max}\rangle\right)$, so that the trivial and nontrivial
Hamiltonians can be perfectly distinguished.
In the case of the two-dimensional version of the ``Grover problem'' with $H_1=
|0\rangle\langle 0|$ and $H_2=|1\rangle\langle 1 |$, this choice for the
driving Hamiltonian actually outperforms the Grover driving term of
Eq.~(\ref{grover_drive}) --- the two Hamiltonians can be distinguished in a
time that is shorter by a factor of $\sqrt{2}$. So while the Grover strategy
is optimal for asymptotically large $N$, it is not actually optimal for $N=2$.
\section{Distinguishing two alternatives in a fixed time}
Let us now suppose that we are to distinguish between two time-independent
Hamiltonians $H_1$ and $H_2$, and that a {\sl fixed duration} $t$ has been
allotted to perform the experiment. Is the driving strategy described above
(in which $-H_2$ is added to the Hamiltonian) always the best possible?
If we have the freedom to add a driving term of our choice, then we may assume
without loss of generality that we are to distinguish the nontrivial
Hamiltonian $H$ from the trivial Hamiltonian $0$. As already noted, if the
largest difference $\Delta E=E_{\rm max}-E_{\rm min}$ of eigenvalues of $H$
satisfies $t\Delta E \ge \pi$, then $H$ can be perfectly distinguished from
$0$; let us therefore suppose that $t \Delta E < \pi$.
If we add a {\sl time-independent} driving term $K$ to the Hamiltonian, and
choose an initial state $|\psi_0\rangle$, then after a time t, we will need to
distinguish the two states
\begin{equation}
\label{two_states}
e^{-i t K}|\psi_0\rangle~, \quad e^{-it(H + K)}|\psi_0\rangle~.
\end{equation}
Two pure states will be more distinguishable when their inner product is
smaller. Therefore, to best distinguish $H+K$ from $K$, we should choose
$|\psi_0\rangle$ to minimize the inner product
\begin{equation}
\left|\langle\psi_0|e^{it K} e^{-it(H + K)}|\psi_0\rangle\right|~.
\end{equation}
If we expand $|\psi_0\rangle$ in terms of the eigenstates $\{|a\rangle\}$ of
$e^{it K} e^{-it(H + K)}$ with eigenvalues $\{e^{-itE_a}\}$,
\begin{equation}
|\psi_0\rangle=\sum_a \alpha_a|a\rangle~,
\end{equation}
this inner product becomes
\begin{equation}
\label{you_tee}
\left|\langle\psi_0|e^{it K} e^{-it(H + K)}|\psi_0\rangle\right|= \left|\sum_a
|\alpha_a|^2 e^{-itE_a}\right|~.
\end{equation}
The right-hand side of Eq.~(\ref{you_tee}) is the modulus of a convex sum of
points on the unit circle. Assuming the modulus is bounded from zero, it
attains its minimum when $|\psi_0\rangle$ is the equally weighted superposition
of the extremal eigenstates of $e^{it K} e^{-it(H + K)}$ -- those whose
eigenvalues are maximally separated on the unit circle.
For $K=0$, the minimum is
$\cos\left( t\Delta E/2\right)$, where $\Delta E$ is the difference of the
maximal and minimal eigenvalues of $H$.
We prove in Appendix A that turning on a nonzero driving term $K$ can never
cause the extremal eigenvalues to separate further, and therefore can never
improve the distinguishability of the two states in
Eq.~{\ref{two_states}.\footnote{That this might be the case was suggested to us
by Chris Fuchs \cite{fuchs_private}.} Therefore, $K=0$ is the optimal driving
term for distinguishing two Hamiltonians. In other words, if we wish to
distinguish between two Hamiltonians $H_1$ and $H_2$, it is always best to turn
on a driving term that precisely cancels one of the two.
The above discussion encompasses the strategy of introducing an ancilla
entangled with the probe (which proved effective for the problem of
distinguishing three or more alternatives). If we wish to distinguish two
Hamiltonians $H_1\otimes I$ and $H_2\otimes I$ that both act trivially on the
ancilla, the optimal driving term exactly cancels one of them ({\it e.g.}, $K=
- H_2\otimes I$), and so it too acts trivially on the ancilla. We derive no
benefit from the ancilla when there are only two alternatives.
Similarly, if we are trying to distinguish only two time-independent signals in
an allotted time, it seems likely there is no advantage to performing a
sequence of weak measurements, and adapting the driving field in response to
the incoming stream of measurement data.
\section{More alternatives: adaptive driving}
Now suppose that there are $N$ possible Hamiltonians ${H_1, H_2,
\ldots, H_N}$. If there is no time limitation, we can distinguish them
perfectly by implementing an adaptive procedure; we make a series of
measurements, modifying our driving term and initial state in response to the
stream of measurement outcomes.
The correct Hamiltonian can be identified by pairwise elimination. First,
assume that either $H_1$ or $H_2$ is the actual Hamiltonian, and apply a
driving term to perfectly
distinguish them, say $H_{D}=-H_1$. After preparing the appropriate initial
state and waiting the appropriate time, we make an orthogonal measurement with
two outcomes --- the result indicates that either $H_1$ or $H_2$ is the
actual Hamiltonian.\footnote{Actually, in a Hilbert space of high dimension, we
can make a more complete measurement that will typically return the result that
neither $H_1$ nor $H_2$ is the actual Hamiltonian.} If the result is $H_1$,
there are two possibilities:
either $H_1$ really is the Hamiltonian, or the assumption that one of $H_1$ or
$H_2$ is the Hamiltonian was wrong. Either way, $H_2$ has been eliminated.
Similarly, if
$H_2$ is found, $H_1$ is eliminated. This procedure can then be repeated,
eliminating one Hamiltonian per measurement, thereby perfectly distinguishing
among the $N$ Hamiltonians in a total of $N-1$ measurements.
This algorithm is quite inefficient, however. The measurement record is $N-1$
bits long, while the information gain is only $\log N$ bits.
\section{Adaptive phase measurement and the semiclassical quantum Fourier
transform}
Far more efficient adaptive procedures can be formulated in some cases.
Consider, for example, a single qubit in a magnetic field of known direction
but unknown magnitude, so that
\begin{equation}
H_\omega= {\omega\over 2}\sigma_3~,
\end{equation}
and let us imagine that the value of the frequency $\omega$ is chosen
equiprobably from among $N=2^n$ equally spaced possible values. Without loss of
generality, we may normalize the field so that the possible values range from 0
to $1-2^{-n}$; then $\omega$ has a binary expansion
\begin{equation}
\omega = .\omega_1 \omega_2 \ldots \omega_n
\end{equation}
that terminates after at most $n$ bits.
The initial state $|\psi_0\rangle={1\over \sqrt{2}}(|0\rangle + |1\rangle)$
evolves in time $t$ to
\begin{equation}
|\psi(t)\rangle_\omega= e^{-itH_\omega}|\psi_0\rangle ={1\over
\sqrt{2}}(|0\rangle + e^{- i \omega t} |1\rangle)
\end{equation}
(up to an overall phase). If we wait for a time $t_n=\pi 2^n$, the
final state is
\begin{equation}
|\psi(t_n)\rangle_\omega={1\over \sqrt{2}}(|0\rangle + e^{-i \pi \omega_n }
|1\rangle) ~.
\end{equation}
Now measurement in the $\{{1\over\sqrt{2}}(|0\rangle \pm |1\rangle\})$ basis
indicates (with certainty)
whether the bit $\omega_n$ is 0 or 1. This outcome divides the set of possible
Hamiltonians in half, providing one bit of classical information.
The set of remaining possible Hamiltonians is still evenly spaced, but it may
have a constant offset, depending on the value of $\omega_n$. However, the
value of
$\omega_n$ is now known, so the offset can be eliminated. Specifically, if we
again prepare $|\psi_0\rangle$ and now evolve for a time $t_{n-1}=\pi
2^{n-1}$, we obtain the final state
\begin{equation}
|\psi(t_{n-1})\rangle_\omega={1\over \sqrt{2}}(|0\rangle + e^{-i \pi \>
(\omega_{n-1} . \omega_n)} |1\rangle)~.
\end{equation}
Since $\omega_n$ is known, we can perform a phase transformation (perhaps by
applying an additional driving magnetic field) to eliminate the phase $e^{-i
\pi \> (. \omega_n)}$;
Measuring again in the $\{{1\over\sqrt{2}}(|0\rangle \pm |1\rangle\})$ basis
determines the value of $\omega_{n-1}$.
By continuing this procedure until all bits of $\omega$ are known, we perfectly
distinguish the $2^n$ possible Hamiltonians in just $n$ measurements. The
procedure is optimal in the sense that we gain one full bit of information
about the Hamiltonian in each measurement.
Up until now we have imagined that the frequency $\omega$ takes one of $2^n$
equally spaced discrete values, but no such restriction is really necessary.
Indeed, what we have described is precisely the implementation of the $n$-qubit
semiclassical quantum Fourier transform as formulated by Griffiths and Niu
\cite{griffiths} (whose relevance to phase estimation was emphasized by Cleve
{\it et al.} \cite{cleve}). Thus the same procedure can be applied to obtain
an estimate of the frequency to $n$-bit precision, even if the frequency is
permitted to take an arbitrary real value in the interval $[0,1)$.
Suppose that we attach to $n$ spins the labels $\{0,1,\ldots,n-2,n-1\}$, and
expose the $k$th spin to the field for time $\pi 2^{k+1}$; we thus prepare the
$n$-qubit state
\begin{equation}
\prod_{k=0}^{n-1} {1\over\sqrt{2}}\left(|0\rangle + e^{-i\pi\omega\cdot
2^{k+1}}|1\rangle\right)={1\over 2^{n/2}}\sum_{y=0}^{2^n-1} e^{-2\pi i \omega
\cdot y}|y\rangle~.
\end{equation}
The adaptive algorithm is equivalent to the quantum Fourier transform followed
by measurement;
hence the $n$-bit measurement outcome $\tilde \omega$ occurs with probability
\begin{equation}
{\rm Prob}_\omega(\tilde \omega) = \left| {1 \over 2^n} \sum_{y=0}^{2^n - 1}
\exp[-2 \pi i y (\omega-\tilde \omega)] \right|^2.
\end{equation}
If $\omega$ really does terminate in $n$ bits, then the outcome $\tilde\omega$
is
guaranteed to be its correct binary expansion. But even if the binary
expansion of $\omega$ does not terminate, the probability that our estimate
$\tilde \omega$ is correct to $n$ bits of precision is still of order
one.\footnote{We might also use the QFT to {\sl compute} eigenvalues of a known
many-body Hamiltonian, rather than {\sl measure} eigenvalues of an unknown one
\cite{lloyd}.}
Of course, to measure the frequency to a precision $\Delta \omega$ of order
$2^{-n}$, we need to expose our probe spins to the unknown Hamiltonian for a
total time $T$ of order $2\pi\cdot 2^{n}$. The accuracy is limited by an
energy-time uncertainty relation of the form $T\Delta\omega\sim 1$.
The semiclassical quantum Fourier transform provides an elegant solution to the
problem of performing an ideal ``phase measurement'' in the Hilbert space of
$n$ qubits. More broadly, any $N$-dimensional Hilbert space with a preferred
basis $\{|k\rangle, ~k=0,1,\dots, N-1\}$ has a complementary basis of {\sl
phase states}
\begin{equation}
|\varphi\rangle={1\over\sqrt{N}} \sum_{k=o}^{N-1}e^{ik\varphi}|k\rangle~,
\end{equation}
with
\begin{equation}
\varphi = 2\pi j/N~,\quad j=0,1,\dots,N-1~.
\end{equation}
For example, the Hilbert space could be the truncated space of a harmonic
oscillator like a mode of the electromagnetic field, with the occupation number
restricted to be less than $N$; then the states $|\varphi\rangle$ are the
``phase squeezed'' states of the oscillator that have minimal phase
uncertainty. Since a POVM in an $N$-dimensional Hilbert space can acquire no
more than $\log N$ bits of information about the preparation of the quantum
state, the phase of an oscillator with occupation number less than $N$ can be
measured to at best $\log N$ bits of accuracy. While it is easy to do an
orthogonal measurement in the occupation number basis with an efficient
photodetector, an orthogonal measurement in the $|\varphi\rangle$ basis is
quite difficult to realize in the laboratory \cite{wiseman}.
But if the standard basis is the computational basis in the $2^n$-dimensional
Hilbert space of $n$ qubits, then an ideal phase measurement is simple to
realize. Since the phase eigenstates are actually not entangled states, we can
carry out the measurement -- {\sl adaptively} -- one qubit at a time.
Note that if we had an arbitrarily powerful quantum computer with an
arbitrarily large amount of quantum memory, then adaptive measurement
strategies might seem superfluous. We could achieve the same effect by
introducing a large ancilla and a driving Hamiltonian that acts on probe and
ancilla, with all measurements postponed to the very end. But the
semiclassical quantum Fourier transform illustrates that adaptive techniques
can reduce the complexity of the quantum information processing required to
perform the measurement. In many cases, an adaptive strategy may be realizable
in practice, while the equivalent unitary strategy is completely infeasible.
\section{Distinguishability and decoherence}
In all of our examples so far, we have ignored noise and decoherence. In
practice, decoherence may compromise our ability to decipher the classical
signal with high confidence. Finding ways to improve measurement accuracy by
effectively coping with decoherence is an important challenge faced by quantum
information theory.
If there is decoherence, our aim is to gain information about the value of a
parameter in a master equation rather than a Hamiltonian. To be concrete,
consider a single qubit governed by an unknown Hamiltonian $H$, and also
subject to decoherence described by the ``depolarizing channel;'' the density
matrix $\rho$ of the qubit obeys the master equation
\begin{equation}
\dot \rho= -i[H,\rho] - \Gamma\left(\rho-{1\over 2} I\right)~,
\end{equation}
where $\Gamma$ is the (known) damping rate.
If we express $\rho$ in terms of the polarization vector $\vec P$,
\begin{equation}
\rho={1\over 2}(I+\vec P\cdot\vec\sigma)~,
\end{equation}
and the Hamiltonian as
\begin{equation}
H={\omega\over 2}~\hat a\cdot \sigma~,
\end{equation}
then the master equation becomes
\begin{equation}
\dot{\vec P}=\omega(\hat a \times \vec P) - \Gamma \vec P~.\
\end{equation}
The polarization precesses uniformly with circular frequency $\omega$ about the
$\hat a$-axis as it contracts with lifetime $\Gamma^{-1}$.
Suppose that we are to distinguish among two possible Hamiltonians, which are
assumed to be equiprobable. If we are able to add a driving term, we may assume
that the two are the trivial Hamiltonian and
\begin{equation}
H={\omega\over 2}~\sigma_3~.
\end{equation}
We choose the initial polarization vector $P_0=(1,0,0)$. Then if the
Hamiltonian is trivial, the polarization contracts as
\begin{equation}
\vec P(t)_{\rm triv}= e^{-\Gamma t}(1,0,0)~,
\end{equation}
while under the nontrivial Hamiltonian it contracts and rotates as
\begin{equation}
\vec P(t)_{\rm nontriv}=e^{-\Gamma t}(\cos \omega t,\sin \omega t, 0)~.
\end{equation}
When is the best time to measure the polarization? We should wait until $\vec
P_{\rm triv}$ and $\vec P_{\rm nontriv}$ point in distinguishable directions,
but if we wait too long, the states will depolarize. The optimal measurement
to distinguish the two is an orthogonal measurement of the polarization along
the axis normal to the bisector of the vectors $\vec P(t)_{\rm triv}$ and $\vec
P(t)_{\rm nontriv}$. At time $t$ the probability that this measurement
identifies the Hamiltonian incorrectly is
\begin{equation}
P_{\rm error}= {1\over 2} - {1\over 2} e^{-\Gamma t}\left|\sin\left({\omega
t\over 2}\right)\right|~.
\end{equation}
This error probability is minimized, and the information gain from the
measurement is maximized, at a time $t$ such that
\begin{equation}
\tan\left({\omega t\over 2}\right)= {\omega\over 2\Gamma}~.
\end{equation}
If $\Gamma/\omega<<1$, this time is close to $\pi/\omega$, the time we would
measure to perfectly distinguish the Hamiltonians in the absence of
decoherence. But if $\Gamma/\omega >>1$, then we should measure after a time
$t\sim \Gamma^{-1}$ comparable to the lifetime.
More generally, consider an ensemble of two density operators $\rho_1$ and
$\rho_2$ with {\it a priori}} probabilities $p_1$ and $p_2$ (where $p_1 +
p_2=1$), and imagine that an unknown state has been drawn from this ensemble. A
procedure for deciding whether the unknown state is $\rho_1$ or $\rho_2$ can be
modeled as a POVM with two outcomes. The two-outcome POVM that minimizes the
probability of making an incorrect decision is a measurement of the orthogonal
projection onto the space spanned by the eigenstates of $p_1\rho_1-p_2 \rho_2$
with positive eigenvalues \cite{helstrom,fuchs_thesis}. The minimal error
probability achieved by this measurement is
\begin{equation}
P_{\rm error}={1\over 2}- {1\over 2}{\rm tr}\left|p_1\rho_1-p_2\rho_2\right|~.
\end{equation}
Correspondingly, if we are to identify an unknown superoperator as one of
$\$_1$ and $\$_2$ (with {\it a priori} probabilities $p_1$ and $p_2$), then the
way to distinguish $\$_1,\$_2$ with minimal probability of error is to choose
our initial state $\rho_0=|\psi_0\rangle\langle \psi_0|$ to
minimize\footnote{We thank Chris Fuchs for a helpful discussion of this point.}
\begin{equation}
\label{super_error}
P_{\rm error}={1\over 2}- {1\over 2}{\rm
tr}\left|\left(p_1\$_1-p_2\$_2\right)\rho_0\right|~.
\end{equation}
In the case of interest to us, the superoperators $\$_1$ and $\$_2$ are
obtained my integrating, for time $t$, master equations with Hamiltonians $H_1$
and $H_2$ respectively. We minimize the error probability in
Eq.~(\ref{super_error}) with respect to $t$ to complete the optimization.
\section{Entanglement and frequency measurement}
Consider again the case in which the Hamiltonian is known to be of the form
\begin{equation}
H_{\omega}={\omega\over 2}~\sigma_3~,
\end{equation}
but where the frequency $\omega$ is unknown. For the moment, let us neglect
decoherence, but suppose that we have been provided with a large number $n$ of
qubits that we may use to perform an experiment to determine $\omega$ in a {\sl
fixed total time} $t$. What is the most effective way to employ our qubits?
Consider two strategies. In the first, we prepare $n$ identical qubits
polarized along the $x$-axis. They precess in the field described by
$H_\omega$ for time $t$, and then the spin along the $x$-axis is measured. Each
spin will be found to be pointing ``up'' with probability
\begin{equation}
P = {1\over 2}(1+ \cos\omega t)
\end{equation}
Because the measurement is repeated many times, we will be able to estimate the
probability $P$ to an accuracy
\begin{equation}
\label{n_qubits}
\Delta P=\sqrt{P(1-P)/n}={|\sin\omega t|\over 2\sqrt{n}}~.
\end{equation}
and so determine the value of $\omega$ to accuracy
\begin{equation}
\label{shot_noise}
\Delta \omega = {\Delta P\over t |dP/d(\omega t)|}={1\over t \sqrt n}~.
\end{equation}
The accuracy improves like $1/\sqrt{n}$ as we increase the number of available
qubits with the time $t$ fixed.
The second strategy is to prepare an entangled ``cat'' state of $n$ ions
\begin{equation}
|\psi_0\rangle = {1\over \sqrt{2}}(|000\dots0\rangle + |111\dots 1\rangle)~.
\end{equation}
The advantage of the entangled state is that it precesses $n$ times faster than
a single qubit; in time $t$ it evolves to
\begin{equation}
|\psi(t)\rangle= {1\over \sqrt{2}}(|000\dots0\rangle + e^{i n\omega t}|111\dots
1\rangle)
\end{equation}
(up to an overall phase).
If we now perform an orthogonal measurement that projects onto the basis
${1\over \sqrt{2}}(|000\dots0\rangle \pm |111\dots 1\rangle)$ ({\it e.g.} a
measurement of the entangled observable
$\sigma_1\otimes\sigma_1\otimes\cdots\otimes\sigma_1$) then we will obtain the
``+'' outcome with probability
\begin{equation}
P={1\over 2}(1+\cos n\omega t)~.
\end{equation}
By this method, $n\omega t$ can be measured to order one accuracy, so that
\begin{equation}\
\label{linear_noise}
\Delta\omega\simeq { 1\over tn}~,
\end{equation}
a more favorable scaling with $n$ than in Eq.~(\ref{shot_noise}).
This idea of exploiting the rapid precession of entangled states to achieve a
precision beyond the shot-noise limit has been proposed in both frequency
measurement \cite{wineland} and optical interferometry \cite{yurke}. (One
realization of this idea is the proposal by Caves \cite{caves} to allow a
squeezed vacuum state to enter the dark port of an interferometer; the
squeezing induces the $n$ photons entering the other port to make correlated
``decisions'' about which arm of the interferometer to follow.)
\section{Entanglement versus decoherence}
In both Eq.~(\ref{shot_noise}) and Eq.~(\ref{linear_noise}), the accuracy of
the frequency measurement improves with the elapsed time $t$ as $1/t$. But so
far we have neglected decoherence. If the single-qubit state decays at a rate
$\Gamma$, then we have seen that the optimal time at which to perform a
measurement will be of order $\Gamma^{-1}$. The entangled strategy will still
be better if we are constrained to perform the measurement in a time
$t<<\Gamma^{-1}$, but further analysis is needed to determine which method is
better if we are free to choose the time $t$ to optimize the accuracy.
In fact, as Huelga {\it et al.} \cite{huelga} have emphasized, an entangled
state is fragile, and its faster precession can be offset by its faster decay
rate. Suppose that two qubits are available, both independently subjected to
the depolarizing channel with decay rate $\Gamma$.
If we prepare the unentangled state, each qubit has the initial pure-state
density matrix
\begin{equation}
\rho_0={1\over 2}(I+\sigma_1)~
\end{equation}
polarized along the $x$-axis, and evolves in time $t$ to
\begin{equation}
\rho(t)={1\over 2}[I+e^{-\Gamma t}(\sigma_1~\cos\omega t+ \sigma_2~\sin\omega
t)]~.
\end{equation}
If we now measure $\sigma_1$, we obtain the $+$ result with probability
\begin{equation}
\label{nonentangle_prob}
\label{single_decohere}
P={\rm tr}\left({1\over 2}(I+\sigma_1)\rho(t)\right)={1\over 2}(1+e^{-\Gamma
t}\cos\omega t)~.
\end{equation}
Now suppose that the initial state is the Bell state $|\phi^+\rangle$ of two
qubits, with density matrix
\begin{equation}
\rho_0={1\over 4}\left(I\otimes I+ \sigma_3\otimes\sigma_3 +
\sigma_1\otimes\sigma_1-\sigma_2\otimes\sigma_2\right)~.
\end{equation}
If both spins precess and depolarize independently, this state evolves to
\begin{eqnarray}
\rho(t)& = & {1\over 4} [I\otimes I+ e^{-2\Gamma t}\big(\sigma_3\otimes\sigma_3
\nonumber\\
& + &\cos 2\omega t(\sigma_1\otimes\sigma_1-\sigma_2\otimes\sigma_2)
\nonumber\\
& + & \sin 2\omega t(\sigma_1\otimes\sigma_2+\sigma_2\otimes\sigma_1)\big)]~;
\end{eqnarray}
if we measure the observable $\sigma_1\otimes\sigma_1$, we find the + outcome
with probability
\begin{eqnarray}
\label{entangle_prob}
P & = &{\rm tr}\left({1\over 2}(I\otimes I
+\sigma_1\otimes\sigma_1)\rho(t)\right)\nonumber \\
& = & {1\over 2}(1+e^{-2\Gamma t}\cos2 \omega t)~.
\end{eqnarray}
Note that Eq.~(\ref{entangle_prob}) has exactly the same functional form as
Eq.~(\ref{nonentangle_prob}), but with $t$ replaced by $2t$. Therefore, the
entangled measurement performed in time $t/2$ collects exactly as much
information about the frequency $\omega$ as the measurement of a single ion
performed in time $t$. If we have two qubits and total time $t$ available, we
can either perform the entangled measurement twice (taking time $t/2$ each
time), or perform measurements on each qubit independently (taking time $t$).
Either way, we obtain two outcomes and collect exactly the same amount of
information on the average.
More generally, suppose that we have $n$ qubits and a total time $T>> 1/
\Gamma$ available. We can use these qubits to perform altogether $nT/t$
independent single-qubit measurements, where each measurement requires time
$t$. Plugging Eq.~(\ref{single_decohere}) into Eq.~(\ref{n_qubits}) and
Eq.~(\ref{shot_noise}) (with $n$ replaced by $nT/t$), and choosing $\cos\omega
t\sim 0$ to optimize the precision, we find that the frequency can be
determined to accuracy
\begin{equation}
\Delta\omega=\left({1\over t}\right)\cdot {e^{\Gamma t}\over
\sqrt{nT/t}}={1\over \sqrt{nT}}\cdot {e^{\Gamma t}\over \sqrt{t}}~.
\end{equation}
This precision is optimized if we choose $\Gamma t=1/2$, where we obtain
\cite{huelga}
\begin{equation}
\Delta\omega = \sqrt{2e\Gamma\over nT}~.
\end{equation}
On the other hand, we could repeat the experiment $T/t$ times using the
$n$-qubit entangled state. Then we would obtain a precision
\begin{equation}
\Delta\omega = \left({1\over nt}\right)\cdot {e^{n\Gamma t}\over
\sqrt{T/t}}={1\over \sqrt{nT}}\cdot {e^{n\Gamma t}\over \sqrt{nt}}~,
\end{equation}
the same function as for uncorrelated qubits, but with $t$ replaced by $nt$.
Thus the optimal precision is the same in both cases, but is attained in the
uncorrelated case by performing experiments that take $n$ times longer than in
the correlated case.
That the entangled states offer no advantage in the determination of $\omega$
was one of the main conclusions of Huelga {\it et al.} \cite{huelga}. A
similar conclusion applies to estimating the difference in path length between
two arms of an interferometer using a specified optical power, if we take into
account losses and optimize with respect to the number of times the light
bounces inside the interferometer before it escapes and is detected.
We would like to make the (rather obvious) point that this conclusion can
change if we adopt a different model of decoherence, and in particular if the
qubits do not decohere independently. As a simple example of correlated
decoherence, consider the case of two qubits with $4 \times 4$ density matrix
$\rho$ evolving according to the master equation
\begin{equation}
\dot \rho = -i[H,\rho] - \Gamma \left(\rho-I/4\right)~.
\end{equation}
This master equation exhibits the analog, in the four-dimensional Hilbert
space, of the uniform contraction of the Bloch sphere described by the
depolarizing channel in the case of a qubit. Because the decoherence picks out
no preferred direction in the Hilbert space (or any preferred tensor-product
decomposition), we call this model ``symmetric decoherence.''
Under this master equation, with both qubits subjected to $H_\omega$ and to
symmetric decoherence, the Bell state $\rho_0=|\phi^+\rangle\langle\phi^+|$
evolves in time $t$ to the state
\begin{eqnarray}
\rho(t)& = & {1\over 4} [I\otimes I+ e^{-\Gamma t}\big(\sigma_3\otimes\sigma_3
\nonumber\\
& + &\cos 2\omega t(\sigma_1\otimes\sigma_1-\sigma_2\otimes\sigma_2)
\nonumber\\
& + & \sin 2\omega t(\sigma_1\otimes\sigma_2+\sigma_2\otimes\sigma_1)\big)]~,
\end{eqnarray}
so that a measurement of $\sigma_1\otimes\sigma_1$ yields the + outcome with
probability
\begin{equation}
\label{ent_symmetric}
P ={1\over 2}(1+e^{-\Gamma t}\cos2 \omega t)~.
\end{equation}
On the other thing, the initial product state
\begin{equation}
\rho_0={1\over 4} (I+\sigma_1)\otimes(I+\sigma_1)
\end{equation}
becomes entangled as a result of symmetric decoherence. Were the Hamiltonian
trivial, it would evolve to
\begin{equation}
\rho(t)={1\over 4} I\otimes I + {1\over 4}e^{-\Gamma t}(\sigma_1\otimes I+
I\otimes\sigma_1 +\sigma_1\otimes\sigma_1)~.
\end{equation}
Including the precession
\begin{equation}
\sigma_1\to \sigma_1 \cos\omega t + \sigma_2\sin\omega t~,
\end{equation}
we obtain
\begin{equation}
\rho(t)={1\over 4} I\otimes I + {1\over 4}e^{-\Gamma t}( \sigma_1\otimes
I~\cos\omega t + \cdots~)~,
\end{equation}
so that measurement of the single-qubit observable $\sigma_1\otimes I$ yields
the + outcome with probability
\begin{equation}
\label{single_symmetric}
P={\rm tr}\left({1\over 2}(I\otimes I +\sigma_1\otimes I)\rho(t)\right)={1\over
2}(1+e^{-\Gamma t}\cos\omega t)~.
\end{equation}
Comparing Eq.~(\ref{single_symmetric}) and Eq.~(\ref{ent_symmetric}), the
important thing to notice is that with symmetric decoherence, entangled states
decay no faster than product states; therefore, we can enjoy the benefit of
entanglement (faster precession) without paying the price (faster decay).
To establish more firmly that entangled strategies outperform nonentangled
strategies in the symmetric decoherence model, we should consider more closely
what are the optimal final measurements for these two types of initial states.
To give the problem a precise information-theoretic formulation, we return to
the problem of distinguishing two cases, the trivial Hamiltonian and
$H_\omega$, which are assumed to be equiprobable. For either the product
initial state or the entangled initial state, we evolve for time $t$, and then
perform the best measurement that distinguishes between evolution governed by
$H_\omega$ and trivial evolution. In both cases, the measurement is permitted
to be an entangled measurement; that is, we optimize with respect to all POVM's
in the four-dimensional Hilbert space.
In either case (initial product state or initial entangled state), we can find
the two-outcome POVM that identifies the Hamiltonian with minimal probability
of error. When there is no decoherence, this POVM (when restricted to the
two-dimensional subspace containing the two pure states to be distinguished) is
the familiar orthogonal measurement that best distinguishes two pure states of
a qubit. In fact, for symmetric decoherence, this same measurement minimizes
the error probability for any value of the damping rate $\Gamma$. It is thus
the two-outcome measurement with the maximal information gain (the measurement
outcome has maximal mutual information with the choice of the Hamiltonian).
Although we don't have a proof, we can make a reasonable guess that, for
symmetric decoherence, this two-outcome measurement has the maximal information
gain of any measurement, including POVM's with more outcomes.
If either initial state evolves for time $t$, and then this optimal POVM is
performed, the error probability can be expressed as
\begin{equation}
P_{\rm error} = {1\over 2} - {1\over 2}e^{-\Gamma
t}\left|\sin\theta(t)\right|~;
\end{equation}
here $\theta(t)$ is the angle between the states --- that is, $\cos \theta(t)$
is the inner product of the evolving and static states, in the limit of no
damping ($\Gamma=0$). For the entangled initial state, we have
\begin{equation}
\theta_{\rm entangled}=\omega t~,
\end{equation}
and for the product initial state, we have
\begin{equation}
\cos\theta_{\rm product}=\cos^2 \left({\omega t\over 2}\right)~.
\end{equation}
Since
\begin{equation}
|\cos \theta_{\rm entangled}| = |\cos\omega t| \le {1\over 2}(1+\cos \omega t)
=|\cos\theta_{\rm product}|
\end{equation}
for $\cos\theta_{\rm entangled}\ge 0$, the error probability achieved by the
entangled initial state is smaller than that achieved by the product state for
$0 < \omega t< \pi/2$, which is sufficient to ensure that the error probability
optimized with respect to $t$ is always smaller in the entangled case for any
nonzero value of $\Gamma$.
Similarly, if we optimize the information gain with respect to $t$, the
entangled strategy has the higher information gain for all $\Gamma>0$. The
improvement in information gain (in bits) achieved using an entangled initial
state rather than a product initial state is plotted in Fig.~1 as a function of
$\Gamma/\omega$. The maximum improvement of about .136 bits occurs for
$\Gamma/\omega\sim .379$.
\begin{figure}[t]
\begin{center}
\leavevmode
\epsfxsize=3in
\epsfbox{entdec.eps}
\end{center}
\caption{Improvement in information gain (in bits) achieved by using an
entangled initial state, as a function of the ratio of decoherence rate
$\Gamma$ to precession frequency $\omega$.}
\end{figure}
We have already seen in \S II that, even in the absence of decoherence, an
entangled strategy may outperform an unentangled strategy if we are trying to
distinguish more than two alternatives. This advantage will persist when
sufficiently weak decoherence is included, whether correlated or uncorrelated.
In that event, since only one member of an entangled pair is exposed to the
unknown Hamiltonian, we may be able to shelter the other member of the pair
from the ravages of the environment, slowing the decay of the state and
strengthening the signal.
\section{Conclusions}
We feel that quantum information theory, having already secured a central
position at the foundations of computer science, will eventually erect bridges
connecting with many subfields of physics. The results reported here (and other
related examples) give strong hints that ideas emerging from the theory of
quantum information and computation are destined to profoundly influence the
experimental physics techniques of the future.
We have only scratched the surface of this important subject. Among the many
issues that deserve further elaboration are the connections between
superoperator distinguishability and superoperator norms, the efficacy of the
quantum Fourier transform in the presence of decoherence, the measurement of
continuous quantum variables, the applications of quantum error correction, and
the detection of time-dependent signals.
\acknowledgments
We thank Constantin Brif, Jon Dowling, Steven van Enk, Jeff Kimble, Alesha
Kitaev, and Kip Thorne for instructive discussions about quantum measurement.
We are especially grateful to Hideo Mabuchi for introducing us to this
fascinating subject, to Chris Fuchs for sharing his insights into state
distinguishability, and to Dave Beckman for discussions on improving the
superdense coding method by applying a driving field. Thanks to Barry Simon for
useful comments on the theorem in Appendix A, and for persuading us that it is
not completely trivial. We also thank C. Woodward for helpful correspondence.
A.~M.~C. and J.~R. received support from Caltech's Summer Undergraduate
Research Fellowship (SURF) program, and A.~M.~C. received a fellowship endowed
by Arthur R. Adams. This work has been supported in part by the Department of
Energy under Grant
No. DE-FG03-92-ER40701, and by DARPA through the Quantum Information and
Computation (QUIC) project administered by the Army Research Office under Grant
No. DAAH04-96-1-0386.
\section*{Appendix A: Fixed-time-driving theorem}
In this appendix, we sketch the proof of the theorem stated in \S V.
For a unitary $N\times N$ matrix $U$, we define ${\rm maxarg}(U)$ to be the
largest argument of an eigenvalue of $U$, where the argument takes values in
the interval $(-\pi,\pi]$. Similarly, ${\rm minarg}(U)$ is the minimum
argument of an eigenvalue of $U$. Our theorem is:\smallskip
{\bf Theorem 1.} {\sl If $H$ and $K$ are finite-dimensional Hermitian matrices,
and $\parallel H\parallel_{\rm sup} <\pi$, then
\begin{eqnarray}
\label{thm1}
{\rm maxarg}\left(e^{iK} e^{-i(H+K)}\right) & \le & {\rm maxarg}(e^{-iH}) ~,\\
\label{thm2}
{\rm minarg}\left(e^{iK} e^{-i(H+K)}\right) & \ge & {\rm minarg}(e^{-iH})~.
\end{eqnarray}
}
To prove the theorem, we begin with:\smallskip
{\bf Lemma 2}. {\sl For unitary $U$ with ${\rm maxarg}(U) \ne \pi$, and
Hermitian $A$,
\begin{eqnarray}
\label{maxlemma2}
& &{\rm maxarg} (Ue^{i\varepsilon A})\le {\rm maxarg}(U) + {\rm
maxarg}(e^{i\varepsilon A})+ O(\varepsilon^2)~,\nonumber\\
& & \\
\label{minlemma2}
& &{\rm minarg} (Ue^{i\varepsilon A})\ge {\rm minarg}(U) + {\rm
minarg}(e^{i\varepsilon A})- O(\varepsilon^2)~.\nonumber\\
& &
\end{eqnarray}
}
\noindent {\sl Proof}: Write $U=e^{iB}$, where $B$ is Hermitian and
\newline $\parallel B \parallel_{\rm sup} < \pi$; then maxarg$(e^{iB})={\rm
max}(B)$, where ${\rm max}(B)$ denotes the maximum eigenvalue of $B$. From the
Baker-Campbell-Hausdorff formula, we have
\begin{equation}
e^{iB}e^{i\varepsilon A}= \exp i\left(B+\varepsilon A + {i\over 2}\varepsilon
[C,B] + O(\varepsilon^2)\right) ~,
\end{equation}
where $C$ is linear in $A$. Then lowest-order eigenvalue perturbation theory
tells us that
\begin{eqnarray}
& &{\rm max}\left(B+\varepsilon A + {i\over 2}\varepsilon
[C,B]\right)\nonumber\\
&=& {\rm max}(B) +\langle \psi|\left(\varepsilon A + {i\over 2}\varepsilon
[C,B]\right)|\psi\rangle +O(\varepsilon^2)\nonumber\\
&=&{\rm max}(B)+\langle\psi|\left(\varepsilon A
\right)|\psi\rangle+O(\varepsilon^2)\nonumber\\
&\le&{\rm max}(B) +{\rm max}(\varepsilon A)+O(\varepsilon^2)
\end{eqnarray}
(where $|\psi\rangle$ is in the eigenspace of $B$ with maximal eigenvalue).
This proves Eq.~(\ref{maxlemma2}). Eq.~(\ref{minlemma2}) is proved similarly.
Note that the condition ${\rm maxarg(U)}\ne\pi$ is necessary so that the
singularity of the maxarg function can be avoided for $\varepsilon$
sufficiently small.
Lemma 2 is all we will need for the proof of Theorem 1. But it is useful to
note that Lemma 2 may be invoked to prove:\smallskip
{\bf Lemma 3}.\footnote{Strangely, we could find only one reference to this
proposition in the literature; it is a special case of Eq.~(8) in
\cite{woodward}.} {\sl For unitary $U_1$ and $U_2$, such that
\begin{eqnarray}
\label{maxargcond}
{\rm maxarg}(U_1) + {\rm maxarg}(U_2) &<& \pi~,\\
\label{minargcond}
{\rm minarg}(U_1) + {\rm minarg}(U_2) &>& -\pi ~,
\end{eqnarray}
we have
\begin{eqnarray}
\label{maxarg}
{\rm maxarg}(U_1 U_2) & \le & {\rm maxarg}(U_1) + {\rm maxarg}(U_2) ~,\\
\label{minarg}
{\rm minarg}(U_1 U_2) & \ge & {\rm minarg}(U_1) + {\rm minarg}(U_2)~.
\end{eqnarray}
}
\noindent{\sl Proof}: We write
\begin{equation}
U_1U_2 = U_1 e^{iA}= U_1 \left(e^{iA/n}\right)^n~,
\end{equation}
where the eigenvalues of A lie in the interval $(-\pi,\pi)$, and apply Lemma 2
repeatedly, obtaining
\begin{eqnarray}
{\rm maxarg}\left(U_1e^{iA}\right)&\le& {\rm maxarg}(U_1)\nonumber\\
&+& n\left[{\rm maxarg}(e^{iA/n})+ O(n^{-2})\right]~.
\end{eqnarray}
Taking the $n\to\infty$ limit proves Eq.~(\ref{maxarg}). Eq.~(\ref{minarg}) is
proved similarly.
Note that because of the conditions Eq.~(\ref{maxargcond}) and
Eq.~(\ref{minargcond}), Lemma 2 can be safely applied $n$ times in succession;
the accumulated maxarg and minarg of the product never approach $\pi$.
To complete the proof of Theorem 1, we invoke the Lie product formula
\begin{equation}
\lim_{n \to \infty} (e^{A/n} e^{B/n})^n = e^{A+B}~,
\end{equation}
to write
\begin{eqnarray}
\label{expexpand}
&&e^{iK}e^{-i(H+K)} = \lim_{n \to \infty} (e^{iK/n})^n (e^{-iH/n}
e^{-iK/n})^n \nonumber \\
& = & \lim_{n \to \infty} e^{iK/n} \cdots e^{iK/n}
e^{-iH/n} e^{-iK/n} \cdots e^{-iH/n} e^{-iK/n}~.
\end{eqnarray}
Since $e^{iK/n} e^{-iH/n} e^{-iK/n}$ and $e^{-iH/n}$ have the same eigenvalues,
Lemma 3 implies that
\begin{eqnarray}
&{\rm maxarg}&(e^{iK/n} e^{-iH/n} e^{-iK/n} e^{-iH/n}) \nonumber\\
& \le & 2 \cdot {\rm maxarg}(e^{-iH/n}) ~.
\end{eqnarray}
Similarly, we have
\begin{eqnarray}
&{\rm maxarg}&\left(e^{iK/n}\left(e^{iK/n} e^{-iH/n} e^{-iK/n} e^{-iH/n}\right)
e^{-iK/n}e^{-iH/n}\right) \nonumber\\
& \le & 3 \cdot {\rm maxarg}(e^{-iH/n}) ~,
\end{eqnarray}
and so on. Hence, applying Lemma 3 altogether $n$ times to the right-hand side
of Eq.~(\ref{expexpand}), we find that
\begin{eqnarray}
{\rm maxarg}\left(e^{iK} \left(e^{-iH/n} e^{-iK/n}\right)^n\right)
& \le & n \cdot {\rm maxarg}\left(e^{-iH/n}\right) \nonumber\\
&= &{\rm maxarg}(e^{-iH}) \\
\end{eqnarray}
Taking the $n\to\infty$ limit completes the proof of Eq.~(\ref{thm1}).
Eq.~(\ref{thm2}) is proved similarly.
The upper bound on $\parallel H\parallel_{\rm sup}$ is a key feature of the
formulation of Theorem 1. This bound ensures that the conditions
Eq.~(\ref{maxargcond}) and Eq.~(\ref{minargcond}) are satisfied each time that
Lemma 3 is invoked in the proof. If $\parallel H\parallel_{\rm sup}$ is too
large, then counterexamples can be constructed.
In any event, for the discussion in \S V, we are interested in the case where
the maximal and minimal eigenvalues of $H$ differ by less than $\pi$, and by
shifting $H$ by a constant we can ensure that $\parallel H\parallel_{\rm sup}<
\pi/2$. Therefore, the theorem enforces the conclusion that if we are to
distinguish a nontrivial Hamiltonian from the trivial Hamiltonian in an
experiment conducted in a fixed elapsed time, turning on a nonzero
time-independent ``driving term'' $K$ provides no advantage.
|
\section{INTRODUCTION}
This is the fifth part of our eight presentations in which we consider
applications of methods from wavelet analysis to nonlinear accelerator
physics problems.
This is a continuation of our results from [1]-[8],
in which we considered the applications of a number of analytical methods from
nonlinear (local) Fourier analysis, or wavelet analysis, to nonlinear
accelerator physics problems
both general and with additional structures (Hamiltonian, symplectic
or quasicomplex), chaotic, quasiclassical, quantum. Wavelet analysis is
a relatively novel set of mathematical methods, which gives us a possibility
to work with well-localized bases in functional spaces and with the
general type of operators (differential, integral, pseudodifferential) in
such bases.
In contrast with parts 1--4 in parts 5--8 we try to take into account
before using power analytical approaches underlying algebraical, geometrical,
topological structures related to kinematical, dynamical and hidden
symmetry of physical problems.
In this paper we consider the applications of discrete wavelet
analysis technique to
maps which come from discretization of continuous nonlinear polynomial
problems in accelerator physics. Our main point is generalization of
wavelet analysis which can be applied for both discrete and continuous
cases. We give explicit multiresolution representation for solutions of
discrete problems, which is correct discretization of our representation
of solutions of the corresponding continuous cases.
In part 2 we consider symplectic and Lagrangian structures for the case of
discretization of flows by corresponding maps and in part 3 construction of
corresponding solutions by applications of
generalized wavelet approach which is based on generalization of
multiresolution analysis
for the case of maps.
\section{Veselov-Marsden Discretization}
Discrete variational principles lead to evolution dynamics analogous to the
Euler-Lagrange equations [9]. Let $Q$ be a configuration space, then a discrete
Lagrangian is a map $L: Q\times Q\to{\bf R}$. usually $L$ is obtained by
approximating the given Lagrangian. For $N\in N_+$ the action sum is the map
$S: Q^{N+1}\to{\bf R}$ defined by
\begin{equation}
S=\sum_{k=0}^{N-1}L(q_{k+1}, q_k),
\end{equation}
where
$q_k\in Q$, $k\ge 0$. The action sum is the discrete analog of the action
integral in continuous case. Extremizing $S$ over $q_1,...,q_{N-1}$ with fixing
$q_0, q_N$ we have the discrete Euler-Lagrange equations (DEL):
\begin{equation}
D_2L(q_{k+1}, q_k)+D_1(q_k, q_{q-1})=0,
\end{equation}
for $k=1,...,N-1$.
Let
\begin{equation}
\Phi: Q\times Q\to Q\times Q
\end{equation}
and
\begin{equation}
\Phi(q_k, q_{k-1})=(q_{k+1}, q_k)
\end{equation}
is a discrete function (map), then we have for DEL:
\begin{equation}
D_2L\circ\Phi+D_1L=0
\end{equation}
or in coordinates $q^i$ on $Q$ we have DEL
\begin{equation}
\frac{\partial L}{\partial q^i_k}\circ\Phi(q_{k+1},q_k)+
\frac{\partial L}{\partial q_{k+1}^i}(q_{k+1},q_k)=0.
\end{equation}
It is very important that the map $\Phi$ exactly preserves the symplectic form
$\omega$:
\begin{equation}
\omega=\frac{\partial^2 L}{\partial q_k^i\partial q_{k+1}^j}(q_{k+1},q_k){\rm
d}q_k^i\wedge{\rm d}q^j_{k+1}
\end{equation}
\section{Generalized Wavelet Approach}
Our approach to solutions of equations (6) is based on applications of general and very efficient methods
developed by A.~Harten [10], who produced
a "General Framework" for multiresolution representation of discrete data.
It is
based on consideration of basic operators, decimation and prediction, which
connect adjacent resolution levels. These operators are constructed from two
basic blocks: the discretization and reconstruction operators. The former
obtains discrete information from a given continuous functions (flows), and
the latter produces an approximation to those functions, from discrete values,
in the same function space to which the original function belongs.
A "new
scale" is defined as the information on a given resolution level which cannot
be predicted from discrete information at lower levels. If the discretization
and reconstruction are local operators, the concept of "new scale" is also
local.
The scale coefficients are directly related to the prediction errors,
and thus to the reconstruction procedure. If scale coefficients are small at a
certain location on a given scale, it means that the reconstruction procedure on
that scale gives a proper approximation of the original function at that
particular location.
This approach may be considered as some generalization of standard wavelet
analysis approach. It allows to consider multiresolution decomposition when
usual approach is impossible ($\delta$-functions case). We
demonstrated the discretization of Dirac function by wavelet packets
on Fig.~1 and Fig.~2.
Let $F$ be a linear space of mappings
\begin{equation}\label{eq:Fin}
F\subset \{f|f: X\to Y\},
\end{equation}
where $X,Y$
are linear spaces. Let also $D_k$ be a linear operator
\begin{eqnarray}
&&D_k: f\to\{v^k\},\quad
v^k=D_kf,\nonumber \\
&&v^k=\{v_i^k\}, \quad v_i^k\in Y.
\end{eqnarray}
This sequence corresponds to $k$ level discretization of $X$.
Let
\begin{equation}\label{eq:Dk}
D_k(F)=V^k={\rm span}\{\eta^k_i\}
\end{equation}
and the coordinates of $v^k\in V^k$ in this basis are
$\hat{v}^k=\{\hat{v}^k_i\}$, $\hat{v}^k\in S^k$:
\begin{equation}
v^k=\sum_i\hat{v}^k_i\eta^k_i,
\end{equation}\label{eq:vk}
$D_k$ is a discretization operator.
Main goal is to design a multiresolution scheme (MR) [10] that applies to all
sequences $s\in S^L$, but corresponds for those sequences $\hat{v}^L\in S^L$,
which are obtained by the discretization (\ref{eq:Fin}).
Since $D_k$ maps $F$ onto $V^k$ then for any $v^k\subset V^k$ there is at least
one $f$ in $F$ such that $D_kf=v^k$. Such correspondence from $f\in F$ to
$v^k\in V^k$ is reconstruction and the corresponding operator is the
reconstruction operator $R_k$:
\begin{equation}
R_k: V_k\to F, \qquad D_kR_k=I_k,
\end{equation}
where $I_k$ is the identity operator in $V^k$ ($R^k$ is right inverse of $D^k$
in $V^k$).
Given a sequence of discretization $\{D_k\}$ and sequence of the corresponding
reconstruction operators $\{R_k\}$, we define the operators $D_k^{k-1}$ and
$P^k_{k-1}$
\begin{eqnarray}
D_k^{k-1}&=&D_{k-1}R_k: V_k\to V_{k-1}\\
P^k_{k-1}&=&D_kR_{k-1}: V_{k-1}\to V_k\nonumber
\end{eqnarray}
If the set ${D_k}$ in nested [10], then
\begin{equation}
D_k^{k-1}P^k_{k-1}=I_{k-1}
\end{equation}
and we have for any $f\in F$ and any $p\in F$ for which the reconstruction
$R_{k-1}$ is exact:
\begin{eqnarray}\label{eq:DP}
D_k^{k-1}(D_kf)&=&D_{k-1}f\\
P^k_{k-1}(D_{k-1}p)&=&D_kp\nonumber
\end{eqnarray}
Let us consider any $v^L\in V^L$, Then there is $f\in F$ such that
\begin{equation}
v^L=D_Lf,
\end{equation}
and it follows from (\ref{eq:DP}) that the process of successive decimation [10]
\begin{equation}
v^{k-1}=D_k^{k-1}v^k, \qquad k=L,...,1
\end{equation}
yields for all $k$
\begin{equation}
v^k=D_kf
\end{equation}
Thus the problem of prediction, which is associated with the corresponding MR
scheme, can be stated as a problem of approximation: knowing $D_{k-1}f$,
$f\in F$, find a "good approximation" for $D_k f$.
It is very important that each space $V^L$ has a multiresolution basis
\begin{equation}
\bar{B}_M=\{\bar{\phi}_i^{0,L}\}_i, \{\{\bar{\psi}^{k,L}_j\}_j\}^L_{k=1}
\end{equation}
and that any $v^L\in V^L$ can be written as
\begin{equation}\label{eq:vL}
v^L=\sum_i\hat{v}_i^0\bar{\phi}_i^{0,L}+\sum^L_{k=1}\sum_j
d_j^k\bar{\psi}_j^{k,L},
\end{equation}
where $\{d_j^k\}$ are the $k$ scale coefficients of the associated MR,
$\{\hat{v}_i^0\}$ is defined by (11) with $k=0$.
If $\{D_k\}$ is a nested sequence of discretization [10] and $\{R_k\}$ is any
corresponding sequence of linear reconstruction operators, then we have from
(\ref{eq:vL}) for $v^L=D_Lf$ applying $R_L$:
\begin{equation}\label{eq:RlDl}
R_LD_Lf=\sum_i \hat{f}^0_i\phi^{0,L}_i+\sum_{k=1}^L\sum_jd_j^k\psi_j^{k,L},
\end{equation}
where
\begin{eqnarray}
&&\phi_i^{0,L}=R_L\bar{\phi}_i^{0,L}\in F,\quad
\psi_j^{k,L}=R_L\bar{\psi}_j^{k,L}\in F,\nonumber\\
&&D_0 f=\sum\hat{f}^0_i\eta^0_i.
\end{eqnarray}
When $L\to\infty$ we have sufficient conditions which ensure that the limiting
process $L\to\infty$ in (\ref{eq:RlDl}, 22) yields a multiresolution basis for $F$.
Then, according to (19), (20) we have very useful representation for solutions
of equations (6) or for different maps construction in the form which are a
counterparts for discrete (difference) cases of constructions from parts 1-4.
\begin{figure}[ht]
\centering
\epsfig{file=tha138a.eps, width=82.5mm, bb=0 200 599 590, clip}
\caption{Wavelet packets.}
\end{figure}
\begin{figure}[ht]
\centering
\epsfig{file=tha138b.eps, width=82.5mm, bb=0 200 599 590, clip}
\caption{The discretization of Dirac function.}
\end{figure}
We are very grateful to M.~Cornacchia (SLAC),
W.~Her\-r\-man\-nsfeldt (SLAC)
Mrs. J.~Kono (LBL) and
M.~Laraneta (UCLA) for
their permanent encouragement.
|
\section{Introduction}
Max Planck discovered the eponymous constant $\hbar$ when studying black body
radiation in 1900 \cite{weav}. He realized immediately that the constants
$\hbar$ and $c$ and $G$ determine a natural scale, now called the Planck
scale, which is easily gotten by dimensional analysis \cite{mtw}. The Planck
distance, time, mass, and energy are
\begin{eqnarray}
L_p \equiv \sqrt{\frac{G \hbar}{c^3}} \simeq 1.6 \times 10^{-35}\text{ m} \, ,
\ \ \ T_p \equiv \frac{L_p}{c} = \sqrt{\frac{G \hbar}{c^5}} \simeq 0.54 \times
10^{-43}\text{ sec} \nonumber \\
M_p \equiv \frac{\hbar}{c L_p} = \sqrt{\frac{\hbar c}{G}} \simeq 2.2 \times
10^{-8} \text{ kg}\, , \ \ \ E_p \equiv \sqrt{\frac{\hbar c^5}{G}} \simeq
2.0 \times 10^9 \text{ J} = 1.2 \times 10^{19} \text{ GeV} \, .\label{planck}
\end{eqnarray}
From its construction the Planck scale should be relevant when the system
considered is quantum mechanical ($\hbar$), involves high velocities and high
energies ($c$), and gravity is important ($G$). One such system is the very
early universe. Another is the collision of elementary particles such as
quarks at about the Planck energy, but to achieve the Planck energy in a
laboratory would probably require an accelerator about the size of a
galaxy. Yet another system in which the Planck scale is relevant is the
cloud of virtual particles surrounding any real particle, since the virtual
particles may in principle have arbitrarily high energies.
Much work has gone into constructing a quantum theory of gravity
appropriate to the Planck scale, but with little practical success. The
only theory thus far that seems to be a plausible candidate is superstring
theory \cite{str}.
Motivation for the present work originated at a talk by John Schwarz at
the Stanford Linear Accelerator Center in 1996, in which he presented a
modified uncertainty principle as a result of superstring theory and scale
inversion symmetry. He asked if, in view of its simplicity, it might be
more general than superstring theory or any particular quantum gravity
theory, and perhaps derivable by simpler means. We hope this work partially
answers that question.
We do not consider any specific quantum gravity theories here, but instead
show that the Heisenberg uncertainty principle is modified when we combine
quantum theory and some basic concepts of gravity. We give four separate
derivations of the modified uncertainty principle: dimensional analysis in
Newtonian theory, an approximate calculation in Newtonian theory,
dimensional analysis in general relativity theory, and an approximate
calculation in general relativity theory \cite{oha}. All four
derivations are heuristic and somewhat rough, as befits a discussion of the
uncertainty principle. Moreover the derivations based on Newtonian theory
should not be taken too seriously since the Newtonian theory is
action-at-a-distance, which is certainly inappropriate for a photon moving
at $c$; our purpose in including the Newtonian derivation is to show that the
modified uncertainty principle appears to follow from rather general
considerations on gravity, in particular that the additional gravitational
term is linear in the energy or momentum of the photon.
A minimum position uncertainty arises immediately from the modified
uncertainty principle, of order of the Planck distance. One might consider
this to be a minimum physically meaningful distance, and thereby question
whether any theory based on a a smaller scale, eg. a spacetime continuum,
really makes operational sense. Additionally, since all measurements
dealing with small distances in particle physics are really large momentum
scattering experiments, the very concept of a spacetime continuum is doubly
suspect. Further speculations of this nature are contained in the
conclusions.
\section{The Uncertainty Principle a la Heisenberg}
Heisenberg in 1923 obtained the uncertainty principle on very general
grounds, based only on general principles of optics and the quantization of
electromagnetic radiation in the form of photons \cite{sch,weav}. We recall his
approach briefly. Consider a wave scattering from an
electron into a microscope and thereby giving a measurement of the position
of the electron. According to optics and intuition, with an
electromagnetic wave of wavelength $\lambda$ we cannot obtain better precision
than
\begin{equation}
\Delta x_H \approx \lambda \, . \label{lamb}
\end{equation}
Such a wave is quantized in the form of photons, each with a momentum
\begin{equation}
p=\frac{h}{\lambda} \, . \label{p}
\end{equation}
In order to interact with the electron an entire photon in the wave must
scatter and thereby impart to the electron a significant part of its
momentum, which produces an uncertainty in the electron momentum of about
$\Delta p \approx p$. Thus we obtain the standard Heisenberg position-momentum
uncertainty relation
\begin{equation}
\Delta x_H \Delta p \approx \lambda\left( \frac{h}{\lambda} \right) \approx h
\approx \hbar \, , \label{up}
\end{equation}
No mention has been made here of the gravitational interaction between the
photon and the electron, which we consider below.
\section{Newtonian Theory, Dimensional Estimate}
We first estimate the effects of gravity in a very rough and heuristic way
using Newtonian gravitational theory, with the assumption that the photon
behaves as a classical particle with an effective mass equal to its energy
divided by $c^2$ \cite{oha}. Suppose the electron is in an experimental
region of characteristic size $L$, inside of which it interacts with the
photon. It will experience an acceleration due to gravity,
\begin{equation}
\ddot{\vec{r}} = - \frac{G (E /c^2)}{r^2} \hat{r} \, , \label{acc}
\end{equation}
where $r$ is the distance between electron and photon. During the
interaction, which occurs in characteristic time $L/c$, the electron will
acquire, due to gravity, a velocity and move a distance, given respectively
by
\begin{equation}
\Delta v \approx \frac{GE}{c^2 r^2} \left( \frac{L}{c} \right) \, , \ \ \ \
\Delta x_G \approx \frac{GE}{c^2 r^2} \left( \frac{L}{c} \right)^2 \, .
\end{equation}
These will be uncertain since the photon scatters electromagnetically from
the electron at some indeterminate time during the interaction. The
electron may be anywhere in the interaction region so the electron-photon
distance should be of order $r \approx L$, which is the only distance scale in
the problem. Since the photon energy is related to the momentum by $E=pc$ we
may also express this as
\begin{equation}
\Delta x_G \approx \frac{Gp}{c^3} \, . \label{xG}
\end{equation}
Noting that the electron momentum uncertainty must be of order of the
photon momentum, and using the Planck length $L_{p}^{2} \equiv G \hbar/ c^3$
as a parameter, we have
\begin{equation}
\Delta x_G \approx \frac{G \Delta p}{c^3} = \left( \frac{G \hbar}{c^3}\right)
\frac{\Delta p}{\hbar}= L_{p}^{2} \frac{\Delta p}{\hbar} \, .\label{xG2}
\end{equation}
This is our main result. We add this uncertainty to the Heisenberg relation
(3) to obtain the modified uncertainty relation
\begin{equation}
\Delta x \approx \frac{\hbar}{\Delta p}+ L_{p}^{2} \frac{\Delta p}{\hbar} \, .
\label{gup}
\end{equation}
We refer to this as the extended uncertainty principle - or more
descriptively as the gravitational uncertainty principle (GUP). Note that
it is invariant under
\begin{equation}
\frac{\Delta p L_p}{\hbar} \longleftrightarrow \frac{\hbar}{\Delta p L_p} \, .
\label{inv}
\end{equation}
That is, it has a kind of momentum inversion symmetry \cite{str}.
\section{Newtonian Theory, Approximate Calculation}
We may also make a more explicit estimate than the above using Newtonian
theory. As before we suppose the electron is in an experimental region of
characteristic size $L$ and interacts with the photon as it crosses the
region. The photon scatters electromagnetically from the electron at some
uncertain time and at some uncertain position inside the experimental
region. Consider first the transverse impulse, that is the motion imparted
to the electron perpendicular to the photon direction. We take the photon
direction to be $x$ and the transverse direction to be $y$ as in figure 2.
The photon passes very rapidly so the electron moves very little in the time it
takes the photon to cross the experimental region, and we may thus take $y
\approx y_0$ that is we use an impulse approximation. The acceleration is then
\begin{equation}
\ddot{y}=\frac{G(E/c^2)}{r^2} \left( \frac{y}{r} \right)= \frac{G (p/c)}{r^2}
\left( \frac{y}{r} \right) \approx \frac{G(p/c)y_0}{(y_{0}^{2} + c^2
t^2)^{3/2}} \, . \label{ay}
\end{equation}
We integrate this to get the transverse velocity impulse,
\begin{equation}
\Delta \dot{y} \approx \frac{2Gp}{c^2 y_0}\left( \frac{cT}{\sqrt{y_{0}^{2}
+ c^2 T^2}}\right) \, ,\label{vy}
\end{equation}
where $T=L/c \ge y_0$ is the characteristic interaction time. We thus have
roughly
\begin{equation}
\Delta \dot{y} \approx \frac{2Gp}{c^2 y_0} \, .\label{vy2}
\end{equation}
Due to this gravitational velocity impulse there will be a change in the
position of the electron, which is intrinsically uncertain, given
approximately by $\Delta y \approx (\Delta \dot{y} / 2) /T$ or
\begin{equation}
\Delta y_G \approx \frac{GpT}{c^2 y_0} \approx \frac{Gp}{c^3} \, ,\label{yG}
\end{equation}
where we have taken $y_0$ to be of order of but less than $L=cT$. This is the
same result (\ref{xG}) as we obtained by somewhat more crude dimensional
arguments in section III.
A similar analysis can be done for motion in the longitudinal direction,
with the result
\begin{equation}
\Delta x_G \approx \frac{Gp}{c^3} \left[ \ln \left( \frac{2L}{x_0} \right) -
1 \right] \, ,
\end{equation}
where $x_0$ is the initial position of the electron. Since $x_0$ must be of
order but less than $L=cT$ and the log is a very slowly varying function we
obtain about the same result as (\ref{yG}) for the longitudinal uncertainty,
\begin{equation}
\Delta x_G \approx \frac{Gp}{c^3} \, . \label{xG3}
\end{equation}
One should be justifiably suspicious of the Newtonian derivations since
Newtonian theory treats the gravitational field in front of the radiation
as action-at-a-distance, whereas the gravitational field actually
propagates at $c$ and cannot extend in front of the radiation. The nature of
the gravitational field will become clear when we discuss the calculation
using linearized general relativity.
\section{General Relativity Theory, Dimensional Estimate}
The arguments in the preceding two sections are only marginally convincing
as heuristic arguments since they are based on action-at-a-distance
Newtonian gravitational theory, with the ad hoc assumption that the energy
of the photon produces a gravitational field. In this section and the
following we give a dimensional estimate and an approximate calculation
based on general relativity theory, free of such drawbacks \cite{abs,mtw,oha}.
The field equations of general relativity are
\begin{equation}
G_{\mu \nu} = - \left( \frac{8 \pi G}{c^4} \right)T_{\mu \nu} \, .
\label{fldeq}
\end{equation}
The left side has the units of inverse distance squared, since it is
constructed from second derivatives and squares of first derivatives of the
metric. Thus on dimensional grounds we may write the left hand side in
terms of deviations of the metric from flat, in schematic order of
magnitude dimensional form, as
\begin{equation}
LHS \approx \frac{\delta g _{\mu \nu}}{L^2} \, , \label{lhs}
\end{equation}
where $\delta g _{\mu \nu}$ denotes the deviation of the metric from
Lorentzian, and $L$ is the same characteristic size as used in sec. III.
Similarly the energy-momentum tensor has the units of an energy density, so
its components must be roughly equal to the photon energy over $L^3$. Thus we
can write the right side of the field equations schematically as
\begin{equation}
RHS \approx \left( \frac{8 \pi G}{c^4} \right) \frac{E}{L^3} \approx \frac{Gp}{
c^3 L^3} \, . \label{rhs}
\end{equation}
Equating the dimensional estimates in (\ref{lhs}) and (\ref{rhs}) we get an
estimate for the deviation of the metric,
\begin{equation}
\delta g_{\mu \nu} \approx \frac{Gp}{c^3 L} \, . \label{dg}
\end{equation}
This deviation corresponds to a fractional uncertainty in all positions in
the region $L$, which we identify with a fractional uncertainty in position,
$\Delta x_G / L$. Thus we have an uncertainty in position due to the
gravitational interaction given by
\begin{equation}
\frac{\Delta x_G }{L} \approx \delta g_{\mu \nu} \approx \frac{Gp}{c^3 L}\, ,
\ \ \ \Delta x_G \approx \frac{Gp}{c^3} \, . \label{xG4}
\end{equation}
As should be expected the characteristic size $L$ has canceled out of the
relation. Finally the uncertainty in momentum of the electron must be
comparable to the photon momentum, $\Delta p \approx p$, and we obtain the same
relation (\ref{xG2}) as before for $\Delta x_G$.
\section{General Relativity Theory, Approximate Calculation}
To make an approximate calculation we use linearized general relativity
theory. From the general energy momentum tensor of the electromagnetic
field it is easy to show that for radiation moving in the $x$ direction the
specific form of the energy- momentum tensor is \cite{tol,abs}
\begin{equation}
T_{mu \nu} = F_{\mu \alpha} F^{\alpha}\; _{\nu} + \frac{1}{4}F_{\alpha \beta}
F^{\alpha \beta} = \rho
\left( \begin{array}{rrrr}
1 \ & -1 \ \ & 0 \ \ & 0 \ \ \\
-1 \ & 1 \ \ & 0 \ \ & 0 \ \ \\
0 \ & 0 \ \ & 0 \ \ & 0 \ \ \\
0 \ & 0 \ \ & 0 \ \ & 0 \ \
\end{array} \right) \, , \label{emt}
\end{equation}
where $\rho=(E^2 + B^2 )/2$ is the energy density of the radiation field, and
may be a function of $x -ct \, , \ y \, , \text{ and } z$, corresponding to a
truncated plane wave. The equations of linearized general relativity theory
follow from (\ref{fldeq}), with the metric taken to be Lorentz plus a small
perturbation, $g _{\mu \nu} = \eta_{\mu \nu} + h_{\mu \nu}$. They are
\begin{equation}
\Box \left [ h_{\mu \nu} - \frac{1}{2} \eta_{\mu \nu} h \right] = - \left(
\frac{8 \pi G}{c^4} \right) T_{\mu \nu} \, , \ \ h \equiv \eta^{\mu \nu} h_{
\mu \nu} = h^{\alpha}_{\alpha}\, ,\label{lfldeq}
\end{equation}
$$
\Box = \frac{\partial^2}{\partial (ct)^2} - \frac{\partial^2}{\partial x^2}
- \frac{\partial^2}{\partial y^2} - \frac{\partial^2}{\partial z^2} \, ,
$$
$$
\text{Lorentz gauge condition } \left[ h_{\mu}^{ \nu} -
\frac{1}{2} \eta_{\mu}^{ \nu} h \right]_{| \nu} =0 \, .
$$
It is straight-forward to solve this system with the energy momentum tensor
given in (\ref{emt}). We are interested only in the inhomogeneous solution, and
the system then reduces to a form involving only one unknown function,
\begin{equation}
h_{\mu \nu} = f(x-ct, y, z)
\left( \begin{array}{rrrr}
1 \ & -1 \ \ & 0 \ \ & 0 \ \ \\
-1 \ & 1 \ \ & 0 \ \ & 0 \ \ \\
0 \ & 0 \ \ & 0 \ \ & 0 \ \ \\
0 \ & 0 \ \ & 0 \ \ & 0 \ \
\end{array} \right) \, , \ \ \
\Box f = - \left( \frac{8 \pi G}{c^4} \right) \rho \, . \label{h}
\end{equation}
For convenience we choose the energy density to be a product,
\begin{equation}
\rho (x-ct,y,z) = \rho_{\parallel}(x-ct) \rho_{\perp}(y,z) \, , \label{rho}
\end{equation}
and it then follows that the metric function $f$ is also a product of the
form $f(x-ct,y,z) = f_{\parallel}(x-ct) f_{\perp}(y,z)$, with
\begin{equation}
\left( \frac{\partial^2}{\partial y^2} + \frac{\partial^2}{\partial z^2}
\right) f_{\perp}(y,z) = \left( \frac{8 \pi G}{c^4} \right) \rho_{\perp}(y,z) \
, , \ \ f_{\parallel}(x-ct) = \rho_{\parallel}(x-ct) \, . \label{wveq}
\end{equation}
Until now the energy density function has been arbitrary. We now choose a
specific function which is convenient for our purposes. For this we take a
cylinder as the envelope of the radiation, inside of which the energy
density oscillates at twice the frequency of the radiation field; the
oscillations however may be ignored since we are interested in an
average uncertainty in position. Thus we take the energy density to be a
constant $\rho_{0}$ inside a cylinder of length $L$ and radius $R$, with $R$
and $L$ comparable in magnitude. Then the solution to (\ref{wveq}), in
cylindrical coordinates $x-ct\, , \ r \, , \ \varphi$, is
\begin{equation}
f = \frac{4G(E/c^2)}{L} g(r) \theta_{L}(x-ct) = \frac{4Gp}{c^3 L} g(r)
\theta_{L}(x-ct) \, , \label{sol}
\end{equation}
$$
g(r) = \left\{ \begin{array}{lr}
r^2/R^2 \, , & r<R \\
1+ \ln(r^2/R^2) \, , & r>R
\end{array} \right\} \, , \ \ \theta_{L}(x-ct) \equiv \theta(x-ct) \theta(ct-x
-L)
$$
Here $E= \pi R^2 L \rho_{0}$ is the total energy of the radiation, and
$\theta_{L}$ is the ``double'' theta function defined above - which is equal to
$0$ ahead of and behind the radiation cylinder. Notice that the gravitational
field does not extend ahead of the radiation field, unlike in the Newtonian
theory. Notice also that the apparent logarithmic divergence of
the external field is not a problem since physical effects involve the
derivative of $f$ which falls off like $1/r$. According to (\ref{h}) and
(\ref{sol}) the metric takes the rather elegant form
\begin{equation}
ds^2=c^2 dt^2 - (dx^2 + dy^2 + dz^2) + f(cdt - dx)^2 \label{g}
\end{equation}
It is clear that since $g(r)$ is of order 1 inside and near the cylinder
of radiation the deviation of the metric from Lorentz is there of order $4Gp /
c^3 L$ and we thereby obtain, using the same idea as in section V,
\begin{equation}
\frac{\Delta x_G }{L} \approx \frac{Gp}{c^3 L} \approx \frac{G \Delta p}{c^3 L}
\, \ \ x_G \approx \frac{4G \Delta p}{c^3} \, . \label{xG5}
\end{equation}
We may also estimate the gravitationally induced motion of the electron
using the above metric and the geodesic equation of motion, to obtain an
alternative derivation of the position uncertainty. The geodesic equation
of motion is
\begin{equation}
\frac{d^2 x^i}{ds^2} + \left\{^{\ i}_{\alpha \beta} \right\}
\frac{dx^{\alpha}}{ds} \frac{dx^{\beta}}{ds}=0 \, , \ \ \alpha,\beta=0,1,2,3
\, , \ \ i=1,2,3 \, . \label{geo}
\end{equation}
For an electron moving reasonably slowly the line element is about $ds=cdt$
and the $\alpha=\beta=0$ term dominates, so the equations of motion are the
usual Newtonian limit equations
\begin{equation}
\frac{d^2 x^i}{dt^2}= - \left\{^{\ i}_{0 0} \right\} c^2 \, . \label{geo2}
\end{equation}
From the metric (\ref{g}) the Christoffel symbols are easily found and we
obtain for motion in the longitudinal $x$ and transverse $r$ directions
\begin{equation}
\frac{d^2 x}{dt^2} = \frac{1}{2} \frac{\partial f}{\partial x} c^2 \, , \ \
\frac{d^2 r}{dt^2} = - \frac{1}{2} \frac{\partial f}{\partial r} c^2 \, .
\label{geo3}
\end{equation}
where we assume motion in the $x$ and $r$ direction.
Longitudinal motion is easy to analyze and rather informative. The
derivative of $f$ gives two delta functions. As the front of the radiation
cylinder passes the electron it first gives the electron a velocity impulse
of
\begin{equation}
\Delta \dot{x} = \frac{2Gp}{c^2 L} g(r) \, , \label{impx}
\end{equation}
and then as the back of the radiation cylinder passes the electron receives
an equal and opposite velocity impulse and stops. In the time of passage it
has moved
\begin{equation}
\Delta x_G = \frac{2Gp}{c^2 L} g(r) T = \frac{2Gp}{c^3 } g(r) \approx
\frac{2Gp}{c^3 } \, , \label{xG6}
\end{equation}
since $g(r)$ is of order 1 in and near the radiation cylinder. Thus we obtain
the same result (\ref{xG2}) as previously. Notice that the gravitational field
of the radiation only acts as it passes over the electron, not before and not
after.
For the transverse motion we differentiate $f$ with respect to $r$ to find
\begin{equation}
\frac{d^2 r}{dt^2} = -\frac{4Gp}{c^2 L}
\left\{ \begin{array}{lr}
r/R^2 \, , & r<R \\
1/r \, , & r>R
\end{array} \right\} \theta_{L}(x-ct)
\end{equation}
so that in the region of the cylinder we have very roughly a velocity
impulse and a corresponding position change $\Delta r \approx (\Delta \dot{r}
/2) T$, given by
\begin{equation}
\Delta \dot{r} = \frac{4Gp}{c^2 LR}T \, , \ \ \Delta r= \frac{2Gp}{c^2 LR}T^2
\approx \frac{2Gp}{c^3}
\end{equation}
That is once again we find that the transverse motion corresponding to an
uncertainty in position due to gravity is the same as in (\ref{xG}) or
(\ref{xG2}).
\section{The Minimal Distance Uncertainty}
The GUP has a remarkable consequence. If the photon momentum and $\Delta p$ are
chosen to be very small then the electron position is imprecise because the
long photon wavelength gives poor resolution. If the photon momentum and
$\Delta p$ are chosen to be very large, then the gravitational field of the
photon makes the electron position very imprecise. Between the two extremes
there is a minimum position uncertainty, which we find from (\ref{gup}) to be
\begin{equation}
\Delta x_{\text{min}} \approx 2\sqrt{\frac{G \hbar}{c^3}} = 2 L_p \, , \
\text{ for } \ \Delta p \approx \sqrt{\frac{\hbar c^3}{G}}\frac{E_p}{c} \,.
\end{equation}
This means that we can never localize the position of a particle such as an
electron to better than about the Planck distance.
Similar analyses of spacetime using the path integral formalism lead to
analogous conclusions; spacetime at small distances and times undergoes
quantum fluctuations, and at the Planck scale the fluctuations are of the
same order as the distances involved \cite{mtw,amel}.
\section{conclusions}
We have shown, using Newtonian and general relativistic gravity, that the
position-momentum uncertainty principle of quantum mechanics is modified by
an additional term. In both theories it is clear that the extra term must
be proportional to the energy or momentum of the photon, so on purely
dimensional grounds the order of magnitude of the extra term is uniquely
determined. As a consequence there is an absolute minimum uncertainty in
the position of any particle such as an electron. Not surprisingly the
minimum is of order of the Planck distance.
In view of the absolute minimum position uncertainty one may plausibly
question whether any theory based on shorter distances, such as a spacetime
continuum, really makes sense. Indeed in light of the fact that laboratory
experiments which probe small distance properties of particles are all high
energy scattering experiments, one might conclude that spacetime at such
small scales may not be a useful concept, and that spacetime at the Planck
scale may not even exist in any meaningful operational sense. Such ideas
are not new and were espoused in the era of S-matrix theory in the 1960s,
that is that the scattering amplitude expressed n terms of input and output
momenta may be the fundamental reality of high energy physics, and not
point-like or string-like particles in a spacetime continuum \cite{sm}.
One might even speculate that the spacetime continuum concept actually
impedes physics in the same way that the concept of an ether impeded
physics in the 19th century. As such, a theoretical structure based entirely
on momenta, such as a modern version of S-matrix theory, might be desirable
and interesting.
\section*{Acknowledgements}
This work was supported by NASA grant NAS 8-39225 to Gravity Probe B. John
Schwarz posed the question studied here in a talk at the Stanford Linear
Accelerator Center in fall 1996, but is of course not responsible for our
interpretation of it. Finally the Gravity Probe B theory group at Stanford
provided many stimulating discussions.
|
\section{Introduction\label{introduction}}
Since about a decade it has been established that dwarf galaxies are
the most numerous type of galaxies in the universe (e.g.\ Mateo
1998)\markcite{mateo98}. Zwicky (1956)\markcite{zwicky56} was among
the first to propose that at least some dwarf galaxies might have been
created from tidal debris strewn about in intergalactic space by
gravitational encounters between massive spiral galaxies. The
relatively nearby M\,81 triplet, for example, is one of the most
stunning examples of how gravitational interactions can redistribute
the neutral atomic gas, originally belonging to individual galaxies,
over a huge volume in the form of bridges and tidal tails (van der
Hulst 1979\markcite{vanderhulst79}, Yun et al.\
1993\markcite{yun:etal93}, Yun, Ho \& Lo\
1994)\markcite{yun:etal94}. The newly born galaxies which may form in
these types of interactions are variously referred to as
`protogalaxies', i.e., those who have not yet started to form stars,
`tidal dwarfs' or `intergalactic molecular complexes'. Recently, they
have also been coined `Phoenix galaxies' (Hernquist
1992\markcite{hernquist92}) since they are presumably born
phoenix--like out of the `ashes' of galaxy--galaxy collisions.
The recent discovery of star forming regions in tidal tails of
interacting galaxies (Mirabel, Lutz \& Maza\
1991\markcite{mirabel:etal91}, Mirabel, Dottori \& Lutz
1992\markcite{mirabel:etal92}; Duc et al.\ 1997\markcite{duc:etal97};
Duc \& Mirabel 1998\markcite{duc:mirabel98}) provides support for the
concept of Phoenix--galaxies. This is corroborated by numerical
simulations which show that tails produced by tidal interactions are
subject to fragmentation (Barnes \& Hernquist
1992\markcite{barnes:hernquist96}; Elmegreen, Kaufman \& Thomasson
1993\markcite{elmegreen:etal93}) which may lead to star formation.
Despite the fact that recent star formation is taking place in tidal
dwarfs, searches for molecular gas in these objects have been rather
unsuccessful thus far. This is surprising as, according to our current
understanding of the star formation process, neutral gas needs to
get dense enough, become self gravitating, and turn largely from
the atomic to a denser molecular phase before star formation can
proceed. Because molecular hydrogen, the main constituent of a
molecular cloud, is difficult to detect at the temperatures and
densities found in molecular clouds, emission from carbon monoxide
(CO) is commonly used as a tracer.
The only example to date of a molecular complex associated with tidal
debris is the one detected by Brouillet et al.\ (1992) within the
M\,81 triplet, close to, but outside the optical image of
M\,81. Although it is presumably located within an \ion{H}{1} arm torn
out of M\,81, no optical emission or star formation appears to be
associated with it (Henkel et al.\
1993\markcite{henkel:etal93}). Since this discovery, a lot of
observational effort has been undertaken to detect similar molecular
complexes in tidal arms of interacting groups of galaxies -- but
without success (e.g., Brouillet, Henkel \& Baudry
1992\markcite{brouillet:etal92}, Smith \& Higdon
1994\markcite{smith:higdon94}). In this paper, we present the
detection of the second molecular complex of this kind.
\section{Observations\label{observations}}
We searched for molecular gas in the tidal arms near NGC\,3077 which
show their presence in observations of the 21\,cm line of neutral
hydrogen (\ion{H}{1}). NGC\,3077 is a member of the nearby
(D=3.2\,Mpc) interacting M\,81 galaxy triplet (see Fig.\ 1, left).
Our CO observations were guided by archival \ion{H}{1} data of
NGC\,3077 (Fig.~1, right, greyscale) obtained with the NRAO Very Large
Array (VLA)\footnote{The National Radio Astronomy Observatory (NRAO)
is operated by Associated Universities, Inc., under cooperative
agreement with the National Science Foundation.} with an improved
angular resolution (10$''$) of a factor of 6 as compared to published
\ion{H}{1} observations (van der Hulst 1979\markcite{vanderhulst79};
Yun et al.\ 1993\markcite{yun:etal93}). The increased resolution
makes it easier to identify regions of high \ion{H}{1} column density
and hence provides better clues as to where to search for molecular
clouds.
The CO observations were carried out in December 1998 with the IRAM
30\, m radio telescope. We observed the $^{12}$CO (${J=1\to0}$) and
(${J=2\! \to \! 1}$) transitions at 115\, GHz and 230\, GHz
simultaneously using two dual channel receivers with a wobbling
subreflector. The wobbler throw was $\pm4'$ in azimuth. The angular
resolution of the telescope is 11$''$ at 230\, GHz and 22$''$ at 115\,
GHz. In total, 13 independent positions were observed within 6 hours
observing time on a $20''$ grid, which is approximately the size of
the telescope beam at 115\,GHz. We used the 1 MHz filter spectrometer
and the autocorrelator spectrometer simultaneously. Spectra for
individual positions were finally degraded to a velocity resolution of
2.6\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\ at both frequencies. The different receiver setups were
combined to one single spectrum per position and transition. Only
linear baselines were removed from the spectra. The final rms ($T_{\rm
mb}$) values for individual positions are listed in Tab.~1.
\begin{deluxetable}{cccccc}
\tablewidth{0pc}
\tablecaption{Gaussian components of the CO ($J=1\!\to\! 0$) spectra.}
\tablehead{
\colhead{$\Delta\alpha$\tablenotemark{a}} &
\colhead{$\Delta\delta$\tablenotemark{a}} &
\colhead{$T_{\rm mb}$} &
\colhead{rms} &
\colhead{$v_{\rm hel}$} &
\colhead{$\Delta v$} \nl
\colhead{$('')$} &
\colhead{$('')$} &
\colhead{(mK)} &
\colhead{(mK)} &
\colhead{(\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi)}&
\colhead{(\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi)}
}
\startdata
0& 60 & -- & 14.0 & $-$ & $-$ \nl
0& 40 & -- & 15.9 & $-$ & $-$ \nl
--20& 20 & -- & 9.6 & $-$ & $-$ \nl
0& 20 & 72.8& 15.1 & $13.9\pm0.9$ & $15.3\pm2.1$ \nl
20& 20 & 47.8& 13.1 & $14.4\pm1.1$ & $12.5\pm3.0$ \nl
--40& 0 & 29.7& 14.8 & $ 9.2\pm2.8$ & $21.7\pm5.5$ \nl
--20& 0 & 43.9& 10.8 & $11.9\pm1.4$ & $23.9\pm3.3$ \nl
0& 0 & 59.4& 8.1 & $15.0\pm0.7$ & $15.1\pm1.6$ \nl
20& 0 & -- & 13.5 & $-$ & $-$ \nl
40& 0 & -- & 13.4 & $-$ & $-$ \nl
--20&--20& -- & 10.9 & $-$ & $-$ \nl
0&--20& 46.0& 14.4 & $14.2\pm1.0$ & $8.5\pm1.9$ \nl
0&--60& -- & 13.9 & $-$ & $-$ \nl
\enddata
\tablenotetext{a}{Positions are offsets relative to $\alpha_{1950}=9^h59^m57.\!^s3$, $\delta_{1950}=68^\circ57'13.\!''8$.}
\label{gauss}
\end{deluxetable}
\begin{figure}
\plotone{walter.fig1.gif.ps}
\caption{ {\em Left:} Optical view of the M\,81--M\,82--NGC\,3077
triplet (taken from the Digitized Sky Survey). A linear scale
(adopting a distance to the triplet of D=3.2\,Mpc) is indicated in the
upper left. The box marks the region around NGC\,3077 which is blown
up on the right. {\em Right:}~Close--up view of the region around
NGC\,3077. The greyscale represents the distribution of the neutral
hydrogen (\ion{H}{1}) as observed with the VLA. The contours
represent the optical image (shown in greyscale on the left). The box
indicates the region where molecular gas was discovered (see Fig.~2
for a close--up view of that region). The linear scale is indicated in
the lower right. Coordinates are given in the B1950.0 epoch.
\label{overview}}
\end{figure}
\section{Results}
Our CO spectra are presented in Figs.\ 2 and 3. Fig.\ 2 gives an
overview of our limited mapping result towards the newly detected
molecular complex. CO ($J=1\to0$) emission was clearly detected from
the central five positions towards the region with the highest
\ion{H}{1} column densities. There is evidence of more emission in
neighboring spectra, but at a lower than 3\,$\sigma$ confidence level.
The properties of the CO ($J=1\to0$) lines as derived from a Gaussian
analysis are listed in Tab.~1. Brouillet et al.\
(1992)\markcite{brouillet:etal92} obtained two spectra close to the
region we mapped. They are ($-2''$, $-73''$) and (240$''$, 16$''$)
offset from our reference position $\alpha_{1950}=9^h59^m57.\!^s3$, $\delta_{1950}=68^\circ57'13.\!''8$. However, Brouillet et
al.\ did not detect CO emission in their spectra at a 1\,$\sigma$--rms
of about 30\, mK in their 5.2\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\ wide channels. Due to the lower
signal-to-noise ratio, emission from the CO ($J=2\to1$) was only
unambiguously detected from our (0,0) position, where we integrated
longest, and for the average spectrum of the five positions where we
detected the lower ($J=1\to0$) transition. For two other positions
where we detected the ($1\to0$) transition there is indication for
($2\to1$) emission also, however with less than 3$\sigma$.
Fig.\ 3 presents a comparison between the spectra of neutral hydrogen
(\ion{H}{1}) and the two lowest rotational CO transitions towards our
reference position. The velocities of the molecular gas in the newly
detected molecular complex are in excellent agreement with the atomic
component. As is also found in other galaxies, the \ion{H}{1} line is
broader than the CO line, indicating that the molecular material is
clumped and present over a smaller range along the line of sight
than the \ion{H}{1}.
The ratio of the two integrated CO lines is (${W_{2\to1}/W_{1\to0})
=0.75\pm0.12}$ for the average of the central five positions; because
the source is clearly extended (s. Fig. 2) no correction for the
different beam areas for the two transitions was made. This ratio is
similar to that found in our Galaxy for quiescent cool (10--20\,K)
molecular clouds (Falgarone et al.\ 1998)\markcite{falgarone:etal98}.
\begin{figure}
\epsscale{0.5}
\plotone{walter.fig2.ps}
\label{spectramap}
\caption{$^{12}$CO (${J=1\!\to\!0}$) spectra obtained towards the
newly detected molecular complex. Positions are offsets relative to
$\alpha_{1950}=9^h59^m57.\!^s3$, $\delta_{1950}=68^\circ57'13.\!''8$. The center of each square corresponds to the location where
each displayed spectrum has been obtained. Note that the spacing
between two individual spectra (20$''$) is approximately the size of
the IRAM 30\,m telescope at 115 GHz (22$''$). The small inserted box
indicates the velocity and temperature scale for each spectrum.}
\end{figure}
\begin{figure}
\epsscale{0.5}
\plotone{walter.fig3.ps}
\label{spectramap}
\caption{$^{12}$CO (${J=1\!\to\!0}$) spectra obtained towards the
newly detected molecular complex. Positions are offsets relative to
$\alpha_{1950}=9^h59^m57.\!^s3$, $\delta_{1950}=68^\circ57'13.\!''8$. The center of each square corresponds to the location where
each displayed spectrum has been obtained. Note that the spacing
between two individual spectra (20$''$) is approximately the size of
the IRAM 30\,m telescope at 115 GHz (22$''$). The small inserted box
indicates the velocity and temperature scale for each spectrum.}
\end{figure}
\section{Discussion\label{discussion}}
\subsection{The properties of the molecular complex}
Because the CO detection coincides with the peak in the \ion{H}{1}
distribution of the tidal gas near NGC\,3077, and strengthened by the
fact that the velocities of the \ion{H}{1} and CO are in excellent
agreement, it is natural to assume that the newly detected molecular
gas is located within the M\,81--M\,82--NGC\,3077 triplet. However, it
is important to further rule out a possible Galactic origin.
The molecular gas is located towards a direction where the Galaxy is
rich in infrared cirrus (de Vries, Heithausen \& Thaddeus\
1987\markcite{devries:etal87}). A significant part of these cirrus
clouds is molecular. However, these clouds are distinct from the
cloud we detected for two reasons. First, molecular cirrus clouds in
this region of the sky have more negative velocities (de Vries et al.\
1987\markcite{devries:etal87}), the closest cirrus cloud is at $v_{\rm
hel}=-4.5$\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\ thus 18.5\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\ off our CO detection. Secondly,
the line width of our molecular cloud ($15\pm1$\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi) is significantly
broader than those of the cirrus clouds ($\Delta v\le4.4$\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi, cf.\
Heithausen 1996\markcite{heithausen96}) and is more similar to that of
giant molecular complexes in our Galaxy. We thus conclude that our
object is not connected to foreground cirrus clouds but extragalactic
in nature.
We therefore assume that the molecular complex is indeed associated
with the tidal debris of NGC\,3077. We adopt a distance of 3.2\, Mpc
(cf.\ Henkel et al.\ 1993\markcite{henkel:etal93}). As shown in Fig.\
1, the region where we detected CO is clearly outside the optical
extent of NGC\,3077. The radius ($R_{25}$) of NGC\,3077 is
$2.\!'7\times2.\!'2$ or 2.5\, kpc $\times$ 2.1\, kpc (de Vaucouleurs
et al.\ 1991\markcite{devaucouleurs:etal91}). Molecular gas of
NGC\,3077 itself is concentrated to its nucleus with a half width at
half maximum (HWHM) of about 0.16\,kpc (Becker, Schilke, \& Henkel
1989)\markcite{becker:etal89}, atomic gas of NGC\,3077 extends further
out (HWHM$\,\approx\,$0.9\,kpc). The projected distance of the newly
detected molecular complex to the center of NGC\,3077 is $3.\!'9$
corresponding to 3.7\, kpc at the distance of NGC\,3077.
If we assume our limited mapping to be complete, molecular gas in the
complex covers an area of 0.4\, kpc$^2$, which corresponds to an
effective radius of about 350\, pc. This is about twice the value
found by Brouillet et al.\ (1992)\markcite{brouillet:etal92} for their
intergalactic complex near M\,81. We note, however, that due to the
low line--intensities observed, we expect beam dilution to play a
major role. This means that the molecular gas in the complex is likely
clumped.
For comparison purpose we calculate a virial mass of the molecular
complex of $M_{\rm vir}\approx 1.7\times10^7$\,M$_{\odot}$, assuming a
constant density throughout the molecular gas. This value is similar
to the object found by Brouillet et al.\
(1992)\markcite{brouillet:etal92} and also similar to the molecular
complex near the nucleus of NGC\,3077 (Becker et al.\
1989\markcite{becker:etal89}). Adopting a Galactic $X$--factor, the
ratio of H$_2$ column density to integrated CO intensity, ${X_{\rm
G}=(1.56\pm0.05)\times10^{20}}$\, cm$^{-2}$\,(K\,\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi)$^{-1}$ (Hunter
et al.\ 1997\markcite{hunter:etal97}), we derive a molecular mass
which is a factor of 20 lower than that based on the assumption of
virialisation. This rather large discrepancy in the mass estimates
could have two reasons: i) the assumption of virialisation does not
hold for this complex; ii) the true $X-$factor in the molecular
complex is comparable to values found in low metallicity galaxies
(Dettmar \& Heithausen 1989\markcite{dettmar:heithausen89}) and in the
Brouillet et al.\ complex. Observations of, e.g., rarer CO isotopomers
are needed to investigate this issue further.
\subsection{The formation of the molecular complex}
The metallicity of interstellar gas can provide important clues as to
the origin of that gas. From the mere fact that we see CO we can
already conclude that the material in the molecular complex near
NGC\,3077 is not primordial, but already pre--processed.
Yun et al.\ (1993)\markcite{yun:etal93} performed detailed numerical
simulations to explain the impressive tidal \ion{H}{1}--structure in
the M\,81 triplet. These simulations suggest that the huge \ion{H}{1}
complex east of NGC\,3077 (total \ion{H}{1} mass: $M_{\rm
HI}=3\times10^8$M$_{\odot}$) where the molecular complex is situated was
formed out of material stripped off the outer parts of NGC\,3077
during the closest encounter with the most massive galaxy of the
triplet, M\,81, about $3\times 10^8$ years ago. Optical studies show
that NGC\,3077's metallicity is around solar (Martin
1997)\markcite{martin97}. This means that even the gas in the
outskirts of NGC\,3077 was already chemically enriched before the
encounter with M\,81 which explains why heavy elements are present.
\subsection{Signs of star formation?}
At this point it is interesting to investigate whether stars have
already formed in the newly detected molecular complex near NGC\,3077
or not. In the case of the intergalactic complex detected by Brouillet
et al.\ (1992)\markcite{brouillet:etal92}, deep optical follow-up
studies showed that no stellar counterpart was
visible\markcite{henkel:etal93}. In our case, an optical study of
this region using the Russian 6\,m telescope has been undertaken by
Karachentsev, Karachentseva \& B\"orngen\
(1985a)\markcite{karachentsev:etal85a} who obtained deep images of an
irregular dwarf object south--east of NGC\,3077 which they named the
`Garland'. As the name indicates, the Garland is a chain of faint
optical knots which occupy an area of $6'\times 4'$ on the sky. Some
diffuse emission is also present in and near the region we mapped.
Karachentsev, Karachenseva \& B\"orngen\
(1985b)\markcite{karachentsev:etal85b} also performed optical
spectroscopy of the four brightest knots in the Garland and determined
a velocity of $(55\pm20)$\ifmmode{{\rm \ts km\ts s}^{-1}}\else{\ts km\ts s$^{-1}$}\fi\ for the knot which is nearest to the
center of our molecular complex ($30''$ to the southwest). The offset
in velocity relative to the complex suggests that the two systems are
probably not associated with each other. The observations therefore
suggest that star formation did not proceed as yet in the newly
detected molecular complex. However, even deeper optical follow up
observations are needed to answer this question unambiguously.
\section{Conclusions\label{conclusions}}
We have established the presence of a giant cold and quiescent
intergalactic molecular complex with no evidence of on--going star
formation. The newly detected molecular complex is embedded in a
large region containing \ion{H}{1} gas and thus has all the
ingredients to form a dwarf galaxy in the future. This object may
therefore be classified as a `Phoenix' galaxy and might resemble the
young dwarf galaxies in the M\,81 group in the future. Our discovery
shows that molecular complexes in tidal arms of interacting galaxies
are not necessarily rare objects --- although they are difficult to
find.
\acknowledgments We thank an anonymous referee for valuable comments
which helped to improve this Letter. FW acknowledges the 'Deutsche
Forschungsgemeinschaft (DFG)' for the award of a stipendium in the
Graduate School "The Magellanic Clouds and other Dwarf Galaxies". We
thank Christian Henkel, Elias Brinks, and Klaas de Boer for fruitful
discussions.
|
\section{Introduction}
\paragraph{}
The black hole solutions of four-dimensional (4d) heterotic string theory
are very interesting from several points of view \cite{RD}. They
give a general description of the whole set of charged black hole
solutions of string theory (and general relativity) in four dimensions.
Heterotic, type $II A$ and type $II B$ strings in four dimensions are
connected
one with the other by a web of dualities \cite{DLR,CA}, so that one
can use the black hole solutions of 4d heterotic string theory as
representative for the whole set of 4d black hole solutions in string theory.
Moreover, using truncated models, it has been realized that there are only
four classes of black holes, which can be described by effective 4d dilaton
gravity models with dilaton coupling $a=\sqrt 3, 1, 1/\sqrt 3, 0$
\cite{DLP,RA}.
The universal character of these solutions has found support both from
the compositeness idea, according to which the $a= 1, 1/\sqrt 3, 0$
solutions can be considered as bound states of the $a=\sqrt 3$
elementary solution \cite {RA} and from the interpretation of them as
intersection of D-branes \cite{PT}.
Each class of 4d heterotic string black holes
is characterized by its spacetime structure, its singularities
and its thermodynamical
behavior.
Particularly remarkable is the behavior of the
$a=0$ black hole (essentially the Reissner-Nordstrom black hole of
general relativity) compared with that of the $a= 1, \sqrt 3, 1/\sqrt 3,$
cases.
The extremal limit of the $a=0$ black hole is a
zero-temperature ground state with non vanishing entropy, whereas for
$a= 1, \sqrt 3, 1/\sqrt 3$ at the extremal limit we have zero temperature
and entropy.
The previous features can be understood using different approaches,
even though a complete and satisfactory picture of
the all subject is still missing. For instance, the existence of
a ground state with the previously described behavior can
be traced back to the presence of a mass gap.
This mass gap can be explained both using string theory \cite{MS1} or just
general features of the 4d effective dilaton gravity theory \cite{HW}.
Until now the various attempts to give a general description and
to clarify the subject have mainly approached
it from ``above'' , i.e using higher dimensional, $d\ge 4$, models.
The recently
proposed AdS$_{d}$ (anti-de Sitter)/ CFT$_{d-1}$ (conformal field theory)
correspondence \cite{Wm} from one side and the discovery of dualities
between four- and two-dimensional black holes \cite{Hy} from the other side,
have changed slightly the perspective.
For $d=2,3$ the AdS/CFT duality has been used to compute the
statistical entropy of 2d and 3d black holes \cite{St,CM}. On the other
hand, it is known that near the horizon the geometry of the 4d, magnetically
charged, $a=0,1,1/{\sqrt 3}$ black hole factorizes as the product of
two spaces of constant curvature, ${\cal M}^{2}\times S^{2}$, where
$S^{2}$ is
the two-sphere and ${\cal M}^{2}$ is 2d Minkowski space for $a=1$ and
${\cal M}^{2}=AdS_{2}$
for $a=0,1/{\sqrt 3}$ \cite{CA}.
It is therefore natural to try to extend the computation of Ref.
\cite {CM} of the
statistical entropy to the 4d case. The hope is that the explanation
of the entropy in terms of microstates will also help to explain the
different behavior of the various black hole solutions.
Motivated by the previous arguments, in this paper we perform a
generic $4d\rightarrow 2d$,
spherical symmetric dimensional reduction of 4d effective heterotic
string theory. We find that the near-horizon, near-extremal behavior
of 4d heterotic string black holes are described by a
class of 2d dilaton gravity models. We show that these 2d models
encode all the relevant information about the 4d theory but in
a much simpler form.
In particular, we show that the
duality group of the 4d theory is realized in the 2d theory in terms
of Weyl
transformations of the metric and
we use the 2d dilaton gravity model
to compute the statistical entropy of the near-extremal 4d,
$a=1/\sqrt3$, black hole.
In Sect. 2 we describe the dimensional reduction $4d\rightarrow 2d$ of
heterotic string theory and the single-scalar field truncation that
produces the effective 2d dilaton gravity models. In Sect. 3 we study
the realization of the 4d duality symmetry in the 2d context, showing
that it corresponds to Weyl transformations of the metric.
In Sect. 4 we use the 2d dilaton gravity model together with the
results of Ref. \cite {CM} to compute the statistical entropy of the
near-extremal, 4d,
$a=1/\sqrt3$, black hole. Finally in Sect. 5 we draw our conclusions.
\section{ Dimensional reduction of 4d heterotic string theory}
\paragraph{}
In the string frame the bosonic action for heterotic string theory
compactified on a six-torus \cite{MS,SE} can be written as follows:
\begin{eqnarray}
A_{H}&=&{1\over 16\pi}\int d^4x\ \sqrt{-g}\ e^{-2 \f}\left\{ R+4(\partial\phi)^{2}-
2 \left[\left( \partial \sigma\right)^2+\left( \partial
\rho\right)^2\right]\right.\nonumber\\
&-& \left.{1\over 4}\left[ e^{-2 \sigma- 2\rho}F_{1}^{2}+e^{-2 \sigma+ 2\rho}F_{2}^{2}+
e^{2 \sigma+2 \rho}F_{3}^{2}
+ e^{2 \sigma-2 \rho}F_{4}^{2}\right]\right\}.
\label{e1}
\end{eqnarray}
In the bosonic action of heterotic string theory
toroidally compactified to $d=4$, we have set to zero the axion fields
and all the $U(1)$ field strengths but four, two Kaluza-Klein fields
$F_{1}, F_{2}$ and two winding modes $F_{3}, F_{4}$. In the
action (\ref{e1})
and throughout this paper we set the
4d Newton constant $G=1$. The scalar fields $\phi,\rho, \sigma$ are related to
the standard definitions of
the string coupling, K\"ahler form and complex structure of the torus,
through the equations
\begin{equation}
e^{-2 \f}={\rm Im}S,\quad e^{-2\sigma}={\rm Im}T,\quad e^{-2\rho}={\rm Im}U.
\label{e2}
\end{equation}
The extremal, Bogomol'ny-Prasad-Sommerfield (BPS), solutions of the action
(\ref{e1}) are given by
(see for instance Ref. \cite{FK})
\begin{eqnarray}\label{e1a}
ds^{2}&=& - \psi_{1} \psi_{3} \, dt^{2}+\left(\chi_{2}\chi_{4}\right)^{-1} (dr^2+r^2d\Omega^2_{2}),
\nonumber\\
e^{4\phi}&=&{\psi_{1}\psi_{3}\over \chi_{2}\chi_{4}},\quad e^{4\sigma}={\psi_{1}\chi_{4}\over \chi_{2}\psi_{3}},\quad
e^{4\rho}={\psi_{1}\chi_{2}\over \psi_{3}\chi_{4}},\nonumber\\
F_{1}&=& \pm d\psi_{1}\wedge dt,\quad
\tilde F_{2}=\pm d\chi_{2} \wedge dt, \quad
\quad F_{3}= \pm d\psi_{3}\wedge dt, \quad
\tilde F_{4}= \pm d\chi_{4} \wedge dt,
\end{eqnarray}
where $(\psi_{1})^{-1}, (\psi_{3})^{-1}, (\chi_{2})^{-1},(\chi_{4})^{-1}$ are harmonic
functions, $d\Omega^2_{2}$ is the metric of the two-sphere and
$\tilde{F}_{2}=
{e^{-2(\phi+\sigma-\rho)}}{^*}F_{2}$,
$\tilde{F}_{4}= {e^{-2(\phi-\sigma+\rho)}}{^*}F_{4}$
($^{*}$ denotes the Hodge dual).
Motivated by the fact that the action (\ref{e1}) admits solutions
that are
the direct product of two two-dimensional spaces of constant
curvature ${\cal M}^{2}\times S^{2 }$ \cite{CA},
we study the general, spherical symmetric, dimensional reduction
$4d \rightarrow 2d$ of the
theory.
Let us first fix our notation. Greek letters from the middle of
the alphabet denote 4d spacetime indices, $\mu ,\nu \ldots =0,1, 2,3$.
Greek letters from the beginning of
the alphabet denote 2d spacetime indices, $\alpha,\beta \ldots =0,1$.
The capital latin letters $I,J..= 2,3 $ label the coordinates of the
internal two-sphere, $ S^{2 }$. The lower-case latin letters
$i,j\ldots=1,2,3,4$ are used to label the four $U(1)$ field
strengths $F_{(i)}$ whereas $a,b\ldots= 1,2,3$ label the
moduli, $\eta_{a}= (\phi,\sigma,\rho)$.
The dimensional reduction of the 4d theory is performed by splitting
the metric and the $U(1)$ field strengths into their 2d parts, using the
ansatz
\begin{eqnarray}
\label{e3}
ds^{2}_{(4)}&=&ds^{2}_{(2)}+ Q^{2}e^{2\psi} d\Omega^2_{2},\nonumber\\
F_{(i) \mu\nu}&=& \left\{ F_{(i) \alpha\beta}, F_{(i) IJ}\right\},
\end{eqnarray}
where $Qe^{\psi}$ is the
radius of the two-sphere, $Q$ is a constant that is related to the $U(1)$
charges (see Eq. (\ref{e6}) below) and the scalar fields $\eta_{a},\psi$
depend only
on the 2d spacetime coordinates.
The 4d field equations together with the ansatz (\ref{e3}) constrain
strongly the form of the fields strengths $F_{(i)}$, we find
\begin{equation}\label{e4}
F_{(i)IJ}= p_{i}\epsilon_{IJ}, \quad F_{(i) \alpha\beta}=
q_{i} e^{2 \eta\cdot b_{i}-2\psi}\epsilon_{\alpha\beta}.
\end{equation}
In the previous equations, and in the following, the dot means
scalar product in the moduli space, $ b_{i}$ are the vectors:
\begin{equation}\label{e5}
b_{1}=(1,1,1), \quad b_{2}=(1,1,-1), \quad b_{3}=(1,-1,-1), \quad
b_{4}=(1,-1,1),
\end{equation}
whereas $ p_{i}, q_{i}$ are respectively the
magnetic and electric charge-vectors that characterize
the particular dimensional reduction.
For sake
of simplicity, we will consider only vectors of the form,
\begin{equation}
\label{e6}
p_{i}=Q\hat{p_{i}}, \qquad q_{i}=Q\hat{q_{i}},
\end{equation}
where $\hat{p_{i}}$ and $\hat{q_{i}}$ are vectors with entries $0$ or $1$.
The form of $ \hat {p_{i}}$ and $\hat { q_{i}}$ determines the 4d
background on which we are performing the dimensional reduction.
In general different charge-vectors will give rise to different 2d
models. The 4d solutions are connected one with the other by O(3,Z) duality
transformations, some of them change the spacetime structure of the
solutions, others leave it invariant \cite {CA}. The most efficient
way to organize the spectrum of the 4d solutions is to use the O(3,Z) duality
symmetries together with the compositeness idea \cite{CA}.
It looks therefore very natural to use the same procedure of Ref.
\cite{CA} to classify
the 2d models deriving from the dimensional reduction of the action
(\ref {e1}). The 4d solutions can be classified in multiplets, labeled
by $N$, of a given number (up to four) of elementary constituent, on
which the duality symmetry $O(3, Z)$ acts in a natural way \cite{CA} .
Moreover, the multiplets with $N=1,2,3,4$ can be put in correspondence
with the
solutions of the single-scalar, single $U(1)$ field strength model
\cite {DLP,RA},
\begin{equation}\label{e6a}
A={1\over 16\pi}\int d^4x\ \sqrt{-g}\ \left\{ R -2(\partial\hat \phi)^{2} -
{1\over 4} e^{-2 a\hat \phi} F^{2}\right\},
\end{equation}
with coupling $a$ given respectively by $a=\sqrt 3, 1, 1/\sqrt 3, 0$.
It turns out that 4d solutions characterized by the same number of
magnetic and electric elementary constituents produce after
dimensional reduction the same 2d model. For this reason the various
dimensional reductions (or equivalently the corresponding 2d models) can be
classified by giving, apart form $N$, the numbers $n,m$ of electric,
respectively, magnetic elementary constituents, with the obvious constrain
$N=n+m$. In this way one obtains 8 different 2d models, which can be
put in correspondence with BPS states of the 4d model
(\ref{e1}),
\centerline{ }
\noindent
\leftline{$N=1$ multiplet, $a=\sqrt 3$,}
\begin{eqnarray}\label{e7}
n&=&1,\quad m=0,\quad \hat q_{i}=(1,0,0,0), \quad \hat p_{i}=0;\nonumber\\
n&=&0,\quad m=1,\quad \hat q_{i}=0,\quad \hat p_{i}=(0,1,0,0).
\end{eqnarray}
\noindent
\leftline{$N=2$ multiplet, $a=1$,}
\begin{eqnarray}\label{e8}
n&=&2,\quad m=0,\quad \hat q_{i}=(1,0,1,0), \quad \hat p_{i}=0;\nonumber\\
n&=&0,\quad m=2,\quad \hat q_{i}=0,\quad \hat p_{i}=(0,1,0,1);\nonumber\\
n&=&1,\quad m=1,\quad \hat q_{i}=(1,0,0,0),\quad \hat p_{i}=(0,0,0,1).
\end{eqnarray}
\noindent
\leftline{$N=3$ multiplet, $a={1\over \sqrt 3}$,}
\begin{eqnarray}\label{e9}
n&=&2,\quad m=1,\quad \hat q_{i}=(1,0,1,0), \quad \hat p_{i}=(0,1,0,0);
\nonumber\\
n&=&1,\quad m=2,\quad \hat q_{i}=(1,0,0,0),\quad \hat p_{i}=(0,1,0,1).
\end{eqnarray}
\noindent
\leftline{$N=4$ multiplet, $a=0$,}
\begin{equation}\label{f1}
n=2,\quad m=2,\quad \hat q_{i}=(1,0,1,0), \quad \hat p_{i}=(0,1,0,1).
\end{equation}
In the previous equations we have given only one representative
element for each model,
characterized by $\hat p_{i}$ and $\hat q_{i}$.
There are, of course, different values of the U(1) charge-vectors that give
rise to the same 2d model. For instance, the model with $N=1, n=1,
m=0$ can be obtained not only with the values of $\hat q_{i}$ given
in Eq. (\ref{e7}), but also when $\hat p_{i}=0 $ and $\hat q_{i}$ has
one entry equal to one, the others being zero.
This degeneracy is related with the duality symmetries of the
theory and will be discussed in detail in the next
section.
In the following we will use $S(N,m)$ to denote a
dimensional reduction associated with a state with $N$ elementary
constituents,
of which $m$ are magnetic ($n=N-m$).
We can now perform explicitly the dimensional reduction of the model
(\ref{e1}) defined by the ansatz (\ref{e3}). The duality symmetry group
$O(3,Z)$ contains off-shell dualities, for this reason it is
convenient to perform the dimensional reduction
at the level of the equation of motion instead of
directly dimensionally reducing the action (\ref{e1}).
Using the ansatz (\ref{e3}) and the equations (\ref{e4}) in the 4d field
equations steming from the action (\ref{e1}), we get the following 2d
field equations,
\begin{eqnarray}\label{f2}
R&+& 4\nabla^{2}\Phi -2 (\partial \Phi)^{2}-2 (\partial \psi)^{2} -2 (\partial \eta)\cdot(\partial
\eta)+\nonumber\\
&+&{1\over 2Q^{4}}e^{2\Phi}\left\{ \sum_{i}q_{i}^{2} e^{2 \eta\cdot b_{i}}-
e^{-4\psi}\sum_{i}p_{i}^{2} e^{-2 \eta\cdot b_{i}}\right\}+ {2\over
Q^{2}}e^{-2\psi}=0,
\end{eqnarray}
\begin{equation}\label{f3}
4\nabla_{\alpha}\left(e^{-2 \F}\nabla^{\alpha}\eta_{a}\right)-{1\over Q^{4}}\left\{ \sum_{i}b_{ia}q_{i}^{2}
e^{2 \eta\cdot b_{i}}-
e^{-4\psi}\sum_{i}b_{ia} p_{i}^{2} e^{-2 \eta\cdot b_{i}}\right\}=0, \quad
a=2,3,
\end{equation}
\begin{equation}\label{f4}
4\nabla_{\alpha}\left(e^{-2 \F}\nabla^{\alpha}\psi\right)+{2\over Q^{4}}e^{-4\p} \sum_{i}
p_{i}^{2} e^{-2 \eta\cdot b_{i}}- {4\over Q^{4}}e^{-2(\psi+\Phi)}=0,
\end{equation}
\begin{eqnarray}\label{f5}
&&\nabla_{a}\nabla_\beta\Phi +\partial_{a}\Phi\partial_\beta\Phi - (\partial_{\alpha} \eta)\cdot(\partial_{\alpha}\eta)-
\partial_{a}\psi\partial_\beta\psi+\nonumber\\
&& - g_{\alpha\beta}\left\{ \nabla^{2}\Phi -{1\over 2} \left[(\partial \Phi)^{2}+
(\partial \psi)^{2} +(\partial \eta)\cdot(\partial \eta)\right]
-{1\over 8Q^{4}}e^{2\Phi}\left( \sum_{i}q_{i}^{2} e^{2 \eta\cdot
b_{i}}+\right.\right.\nonumber\\
&&\left.\left.e^{-4\psi}\sum_{i}p_{i}^{2} e^{-2 \eta\cdot b_{i}}\right)+ {1\over
2 Q^{2}}e^{-2\psi}\right\}=0,
\end{eqnarray}
where the curvature $R$ and all the differential operators are
2d quantities, the field $\Phi$ is the 2d dilaton related, as usual,
to the 4d
dilaton $\phi$ by
\begin{equation}\label{f6}
\phi=\Phi-\psi,
\end{equation}
and the vector $\eta_{a}$ is now defined in terms of $\Phi$,
$\eta_{a}=(\Phi,\sigma,\rho)$.
The field equations (\ref{f2})-(\ref{f5}) are rather complicated.
They assume a much simpler form by considering a consistent,
single-scalar
field truncation that reduces them to those of 2d gravity coupled to the
dilaton $\Phi$. These models have been widely investigated in recent years
under the name of 2d dilaton gravity.
The single-scalar field truncation is obtained using an ansatz
expressing the fields $\psi,\sigma,\rho$ in terms of the dilaton $\Phi$ in a way
that is consistent with the field equations (\ref{f2})-(\ref{f5}),
\begin{eqnarray}\label{f7}
e^{2\psi}&=&A^{2} \exp\left( {{4(m-2)\over a^{2}N}\,\Phi}\right),\nonumber\\
e^{2\sigma}&=&A^{-1} \exp\left( {{2C_{\sigma}\over a^{2}N}\,\Phi}\right) ,\nonumber\\
e^{2\rho}&=&A^{-1} \exp\left( {{2C_{\rho}\over a^{2}N}\,\Phi}\right),
\end{eqnarray}
where $A=\sqrt{3/2}$ for $N=1$, $A=1$ for $N=2,3,4$.
$C_{\sigma}=-1$ for the states $S(1,1), S(2,1),$ $S(3,2)$;
$C_{\sigma}=0$ for $S(2,2), S(2,0), S(4,2)$;
$C_{\sigma}=1$ for $S(1,0), S(3,1)$.
$C_{\rho}=-1$ for the states $S(1,1), S(1,0), S(3,2),$ $ S(3,1)$;
$C_{\rho}=0$ for $S(2,2), S(2,0), S(2,1),$ $ S(4,2)$.
After inserting Eqs. (\ref{e7})-(\ref{f1}) and (\ref{f7}) into the field
equations (\ref{f2})-(\ref{f5}), we find that, for $N=1,2,3$, they are
equivalent to those
derived from a class of 2d dilaton gravity models whose action has the
form,
\begin{equation} \label{f8}
A_{2}={1\over 2}\int d^2x\ \sqrt{-g}\ e^{-2 \F}\left[ R+\Omega (\partial\Phi)^{2}+\l^2 V(\Phi)\right],
\end{equation}
where $\l^2=1/Q^{2}$ for $N=1,2$, and $\l^2=1/(2Q^{2})$ for $N=3$.
The kinetic coefficient $\Omega$ and the dilaton potential $V$ are
completely determined in terms of $n, m,a, N,$
\begin{equation}\label{f8a}
\Omega = 8\left( {1-n\over a^{2 }N}\right), \qquad V(\Phi)= \exp\left[ 4\left(
{2-m\over a^{2}N}\right)\Phi\right].
\end{equation}
For $N=4$ the field equations (\ref{f2})-(\ref{f5}) become equivalent to those
derived from the action,
\begin{equation}\label{f9}
A={1\over 2}\int d^2x\ \sqrt{-g}\ e^{-2 \F}\left[ R+4(\partial\Phi)^{2}- {1\over 2} F^{2}+\l^2 \right],
\end{equation}
which describes the heterotic string in 2d target-space \cite{NY}.
The class of dilaton gravity models defined by the action (\ref{f8})
contains, as particular cases, models that have been already investigated
in the past. For $N=2, m=2$ the action describes the
Callan-Giddings-Harvey-Strominger (CGHS) model \cite{CGHS}. For $N=2,
m=1$ we get
the Weyl rescaled CGHS model investigate in Ref. \cite {CM3}.
Finally, for $N=3,
m=2$ we obtain the Jackiw-Teitelboim (JT) model \cite{JT}.
In general the 2d dilaton gravity model (\ref{f8}) give a near-horizon
description of the, near-extremal, 4d black hole solution of the action
(\ref{e1}).
This near-horizon description corresponds to
setting to zero the constant term in the harmonic functions
appearing in Eq. (\ref{e1a}).
On the other hand, one can easily verify that this is exactly the way to
make the ansatz
(\ref {f7}) consistent the 4d extremal solutions (\ref{e1a}).
\section{Dualities and Weyl transformations}
\paragraph{}
In the canonical frame $g_{C}= e^{-2\phi}g_{S}$, the 4d field equations
following from the action (\ref{e1}) are invariant under the action
the
$O(3,Z)$ duality group
that acts on the moduli $\phi,\sigma,\rho$ and on the $U(1)$ field strengths but
leaves the 4d metric unchanged \cite{CA}. The $O(3,Z)$ duality group
can be generated using three $S-T-U$ duality transformations
$\tau_{S},\tau_{T},\tau_{U}$ together with the permutation group $P_{3}$
acting on the moduli $\phi,\sigma,\rho$ and on the $U(1)$ field
strengths
$F_{i}$ (see Ref.\cite{CA}). After
passing to the string metric $g_{S}$ and performing the
dimensional reduction described in the previous section,
the duality group $O(3,Z)$ becomes a symmetry of the 2d field
equations (\ref{f2})-(\ref{f5}). Because the 4d action (\ref{e1})
is written in
terms of the
string metric the duality transformation will act also on the 2d metric
$g_{\alpha\beta}$ and, owing to Eq. (\ref{e4}), also on the
charge-vectors $p_{i}, q_{i}$. One can easily verify that the O(3,Z)
duality
transformations leave invariant the 2d dilaton field $\Phi$, whereas
in terms of the remaining 2d fields $\psi,\rho,\sigma,g_{\alpha\beta}$ and of
the charge vectors $p_{i}, q_{i}$
it is realized as follows:
\begin{eqnarray}\label{g1}
\tau_{S}&:& \psi\rightarrow -\psi -2\Phi,\quad g_{\alpha\beta}\rightarrow e^{-4 (\Phi+\psi)}g_{\alpha\beta},
\quad q_{1}\rightarrow p_{3},\quad
q_{3}\rightarrow p_{1},\quad q _{2}\rightarrow p_{4},\nonumber\\
& & q_{4}\rightarrow p_{2};\nonumber\\
\tau_{T}&:& \sigma\rightarrow -\sigma,\quad q_{1} \leftrightarrow q_{4},\quad
q_{2}\leftrightarrow q_{3};\nonumber\\
\tau_{U}&:& \rho\rightarrow -\rho,\quad q_{1}\leftrightarrow q_{2},\quad
q_{3}\leftrightarrow q_{4};
\end{eqnarray}
\begin{eqnarray}\label{g2}
P_{3}: & &\sigma \leftrightarrow \rho,\quad q_{2}\leftrightarrow q_{4};
\nonumber\\
& &\psi\rightarrow \sigma-\Phi,\quad \sigma\rightarrow \psi +\Phi,\quad g_{\alpha\beta}\rightarrow e^{2 ( \sigma -\Phi-\psi)}g_{\alpha\beta},
\quad q_{3}\rightarrow p_{4},\quad q_{4}\rightarrow p_{3};
\nonumber\\
& &\psi\rightarrow \sigma-\Phi,\quad\sigma\rightarrow \rho,\quad \rho\rightarrow \Phi+\psi,\quad g_{\alpha\beta}\rightarrow
e^{2 ( \sigma -\Phi-\psi)}g_{\alpha\beta},
\quad q_{4}\rightarrow q_{2}, \nonumber\\
& &q_{2}\rightarrow p_{3},\quad
q_{3}\rightarrow p_{4};
\nonumber\\
& &\psi\rightarrow \rho-\Phi,\quad \rho\rightarrow \Phi+\psi,\quad g_{\alpha\beta}\rightarrow e^{2 ( \rho -\Phi-\psi)}g_{\alpha\beta},
\quad q_{2}\rightarrow p_{3}, \quad
q_{3}\rightarrow p_{2};
\nonumber\\
& &\psi\rightarrow \rho-\Phi,\quad\sigma\rightarrow\Phi+\psi,\quad \rho\rightarrow \sigma,\quad g_{\alpha\beta}\rightarrow
e^{2 ( \rho -\Phi-\psi)}g_{\alpha\beta},
\quad q_{2}\rightarrow q_{4}, \nonumber\\
& &q_{3}\rightarrow p_{2},\quad
q_{4}\rightarrow p_{3}.
\end{eqnarray}
The previous equations describe the action of the duality group
on electric states ($p_{i}=0$), the action on magnetic states
($q_{i}=0$) can be easily obtained from Eqs. (\ref{g1}),(\ref{g2}),
by interchanging $q_{i}\leftrightarrow p_{i}$.
The duality group generated by the transformations (\ref{g1}),(\ref{g2})
becomes extremely simple once we perform the single-scalar field
truncation described in Sect. 2. One can easily verify that the
transformations $\tau_{T},\tau_{U}$ and the first transformation in Eq.
(\ref{g2}) change the ansatz (\ref{f7}) but not the resulting 2d
action (\ref{f8}), (\ref{f9}). This fact has a natural explanation.
$\tau_{T},\tau_{U}$
and the first duality in Eq. (\ref{g2})
do not change the number $n$ of electric (or magnetic $m$) elementary
constituents of a state, so that they cannot change the 2d action
because the latter is parametrized in terms of $n$ and $m$ only.
On the other hand, using Eqs. (\ref{f7}) into Eqs.
(\ref{g1}),(\ref{g2}), one finds that the
remaining duality transformations (in particular the $\tau_{S}$ duality)
act on the 2d dilaton gravity models
(\ref{f8}) and (\ref{f9}) as Weyl transformations of the 2d metric,
\begin{equation}
\label{g3}
g_{\alpha\beta}\rightarrow e^{P\Phi} g_{\alpha\beta}, \quad P={4\over a^{2}N} \left( m-m'\right),
\end{equation}
which map one into the other models in Eqs. (\ref{f8}), (\ref{f9})
with the same value of $N$
but with a number $m$ and $m'$ of magnetic elementary constituents.
Acting at $N$ fixed, the Weyl transformations (\ref{g3}) connect one
with the other models within a given multiplet.
For instance, taking $N=2$ we have three models:
$S(2,2)$ ($\Omega = 4$, $V(\phi)=1$, the CGHS model), $S(2,1)$ ($\Omega = 0$,
$V(\phi)=\exp (2\Phi)$, the Weyl-rescaled CGHS model of Ref. \cite{CM3} and
$S(2,0)$ ($\Omega = -4$,
$V(\phi)=\exp (4\Phi)$). These models are obtained one from the other using
the Weyl transformations (\ref{g3}). $g_{\alpha\beta}\rightarrow e^{2\Phi} g_{\alpha\beta}$,
maps $S(2,2)\rightarrow S(2,1)$ and $S(2,1)\rightarrow S(2,0)$ whereas
$g_{\alpha\beta}\rightarrow e^{4\Phi} g_{\alpha\beta}$ maps $S(2,2)\rightarrow S(2,0)$.
The $\tau_{S}$ duality is essentially an electro/magnetic duality, so
that the strong/weak coupling duality of the 4d theory
(\ref{e1}) is translated, after dimensional reduction to two
dimensions, at least for the 4d model under
consideration, into a
dilaton-dependent Weyl
rescaling of the 2d metric.
It has been shown that dilaton-dependent Weyl transformations leave
invariant the physical parameters (the mass and for 2d black hole
solutions, the Hawking temperature and radiation flux) associated with
the solutions of 2d dilaton gravity \cite{CA2}.
Also other physical parameters that can be expressed in terms of the
mass and the temperature (e.g the entropy) are invariant under such
transformations.
The dimensional reduction
seems to wash out most of the information about the magnetic or
electric character of the 4d solution. At the 2d level the relevant
information is encoded in the number $N$, the number of elementary
constituents, the actual number of magnetic and electric elementary
constituents being irrelevant for the physical parameters of the
solution.
At the level of the 4d theory this implies the equivalence of electrically and
magnetically
charged configurations, as long as only excitations near
extremality are concerned. Moreover, in the 2d context, the duality
implies the equivalence of spacetime structure that behave rather
differently. 2d spacetimes of constant curvature ( e.g the solutions
of the 2d model $S(3,2)$) are dual (i.e connected by Weyl
transformations) to spacetimes with singularities (e.g the solutions
of the 2d model $S(3,1)$).
Presently, we do not know if this is a peculiarity of the 4d
heterotic string theory (\ref{e1}) or a general feature of 4d effective
string theory. However, our results indicate that the dimensionally
reduced 2d theory takes care only of the relevant physical properties of
the 4d model and can therefore be used to give a universal
classification of the near-extremal behavior of the 4d black hole
solutions of string theory.
\section{ Statistical entropy of the a=1/${\bf\sqrt 3}$ 4d black hole}
\paragraph{}
The results of the previous sections together with those of Ref. \cite{CM}
can be used to calculate,
microscopically, the entropy of the near-extremal 4d, $a=1/\sqrt3$, black hole.
The near-extremal, near-horizon behavior of this
black hole is described by the 2d model of Eq. (\ref{f8}) with
$N=3, m=2$, i.e by the JT dilaton gravity model.
For the JT black hole a derivation of
the statistical entropy has been given in Ref. \cite{CM}, one can, therefore,
use it to compute, microscopically, the entropy of the
4d, $a=1/\sqrt3$, black hole.
For $a=1/\sqrt3$ the single-scalar field model of Eq.(\ref {e6a})
takes the form,
\begin{equation}\label{h1}
A={1\over 16\pi}\int d^4x\ \sqrt{-g}\ \left\{ R-2(\partial\hat \phi)^{2}-
{1\over 4} e^{-{2\over \sqrt 3}\hat \phi}F^{2}\right\},
\end{equation}
where the scalar field $\hat \phi$ is connected to the 4d dilaton $\phi$
trough $ \hat \phi= \sqrt 3 \,\phi$.
As mentioned above, this model arises as single-scalar field, single $U(1)$
field strength truncation of the $N=3$ composite solutions of the action
(\ref{e1}) and as compactification of the five-dimensional (5d)
Einstein-Maxwell theory.
The general (non extremal) black hole solution of the model has the
form \cite{DLP1},
\begin{eqnarray}\label{h2}
ds^{2}&=&- H^{-{3\over2}}\left( 1-{\mu\over r}\right) dt^{2}+
H^{3\over2}\left( 1-{\mu\over r}\right)^{-1} dr^2+ H^{3\over2} r^2 d\Omega^2_{2},
\nonumber\\
e^{2\phi}&=& H^{1\over 2},\quad H=1+{\mu \sinh ^{2} \alpha \over r},
\end{eqnarray}
where for simplicity we have set the constant mode of the 4d
dilaton $\phi_{0}=0$.
The integration constants $\mu$ and $\alpha$ are related to the mass and
charge of the solution by
\begin{equation}\label{h3}
M= {1\over 2} \mu \left( 1+{3\over 2}\sinh ^{2} \alpha\right),
\quad Q={1\over 2} \sinh 2 \alpha.
\end{equation}
To be more precise, $Q$ is the (common) charge of the three $U(1)$
fields in the action (\ref{e1}). $Q$ is related to the charge $\hat Q$
of the single $U(1)$ field appearing in the action (\ref{h1}) by $Q=
(2/\sqrt 3)\hat Q $.
Using the area law we find that the Bekenstein-Hawking entropy of the hole
is given by
\begin{equation}\label{h4}
S={{\cal A} \over 4}= \sqrt \mu\left( \mu+ \mu\sinh ^{2} \alpha\right)^{3/2},
\end{equation}
where ${\cal A}$ is the area of the event horizon.
The extremal black hole is obtained in the limit $\mu\rightarrow 0,\, \alpha \rightarrow
\infty$, keeping $\mu \sinh ^{2} \alpha=Q$. In this limit the solution is
given by Eq. (\ref {h2}) with $\mu=0$ and $H=(1+ Q/r)$ whereas the mass
is $M_{ex}= {3\over 4}Q$.
Let us now consider small excitations near the extremal solution, the
entropy of these configurations is given by
\begin{equation}\label {h4a}
S_{(4)}= \pi Q^{3/2} \sqrt
\mu
+\ord{\mu^{3/2}}.
\end{equation}
We know that, when expressed in terms of the string metric, the near-horizon,
near-extremal, magnetically charged, black hole solution (\ref{h2})
factorizes as ${\cal M}^{2}\times S^{2}$, where ${\cal M}^{2}$ is the
solution of the 2d dilaton gravity model (\ref{f8}) with $N=3,\, m=2,
a=1/\sqrt 3$ ,
\begin{equation}\label{h5}
A={1\over 2}\int d^2x\ \sqrt{-g}\ e^{-2 \F}\left( R+2\l^2 \right).
\end{equation}
(We have rescaled $\l^2 \rightarrow 2\l^2$ in order to mach the conventions of
Ref. \cite{CM}).
The dilaton gravity model (\ref{h5}) admits solutions that can be
interpreted as black holes in 2d AdS space \cite{CM2},
\begin{equation}\label{h6}
ds^2=-(\lambda^2x^2-b^2)dt^2+(\lambda^2x^2-b^2)^{-1}dx^2,\qquad e^{-2 \F}=e^{-2\Phi_{0}} \lambda x.
\end{equation}
The mass $M_{(2)}$ and the entropy $S_{(2)}$ of the 2d black hole are
given in terms of the integration constants $b, \Phi_{0}$, by
\begin{equation}\label{h7}
M_{(2)}={1\over 2} e^{-2\Phi_{0}} b^2\lambda, \quad S_{(2)}=4\pi \sqrt{ e^{-2\Phi_{0}}
M_{(2)}\over 2\lambda}.
\end{equation}
If the 2d model (\ref{h5}) has to describe the near-extremal,
near-horizon regime
of the 4d black hole solutions (\ref{h2}) then the 2d expression for
the entropy (\ref{h7}) should match the leading order of the
corresponding 4d quantity in Eq. (\ref{h4a}).
To show this, we first write the 4d solution (\ref{h2}) in terms of the string
metric, we expand the solution near extremality and near the horizon.
After some manipulations
we get
\begin{eqnarray}\label{h7a}
ds^{2}&=&-\left(\l^2 x^{2} -\lambda \mu \right) dt^{2}+
\left(\l^2 x^{2} -\lambda \mu \right)^{-1} dx^2+ e^{2\phi_{0}}Q^{2} d\Omega^2_{2},
\nonumber\\
e^{-2\phi}&=&\sqrt 2 \lambda x.
\end{eqnarray}
where $\lambda = 1/(2Q)$.
As expected the 4d solution factorizes as the product of a 2d spacetime
and a two-sphere of constant radius.
Taking into account that the dimensional reduction $4d\rightarrow 2d$ implies
the following relation between the 2d dilaton $\Phi$ and the 4d one
$\phi$:
\begin{equation}\label{h8}
e^{-2 \F}= {1\over 2} Q^{2} e^{-2 \f},
\end{equation}
and comparing Eq. (\ref{h7a}) with Eq. (\ref{h6}), we get
\begin{equation}\label{h8a}
\mu={2M_{(2)}\over \l^2} e^{2\Phi_{0}},\quad e^{-2\Phi_{0}}={\sqrt 2\over
8\l^2}.
\end{equation}
Using Eqs. (\ref{h8a}) into the expression (\ref{h4a}) for
the 4d entropy and taking into account only the leading
term, we obtain a complete
agreement with the 2d results, i.e $S_{(4)}=S_{(2)}$.
Until now we have considered only 4d, magnetically charged, solutions,
i.e the state $S(3,2)$. The 4d electrically charged solution $S(3,1)$
does not factorize as
direct product of two 2d spaces. Near extremality it is described by the
2d model with $N=3, m=1$, which is dual to the model (\ref {h5}).
Because the 2d entropy does not change under Weyl rescaling of the
metric, it follows that the Eqs. (\ref{h7}) and the equality
$S_{(4)}=S_{(2)}$ hold also for excitations
near 4d extremal, electrically charged, solutions.
We have shown that the semiclassical dynamics of small excitations near
extremality of the 4d black hole can be described by the 2d model
(\ref{h5}) and that at the leading order, the 2d and 4d thermodynamical
entropy is the same.
One can therefore use the results of Ref.
\cite{CM} as an indirect calculation of the statistical entropy of the 4d
black hole (\ref{h2}) in the near-extremal regime.
In Ref. \cite{CM} we have found a mismatch of a $\sqrt 2$ factor between the
thermodynamical and the statistical entropy of the 2d black hole.
This implies that also the
statistical entropy of the 4d $a=1/\sqrt 3$ black hole agrees only up
to a $\sqrt 2$ factor with the thermodynamical result. A simple
explanation of this $\sqrt 2$ factor could be found when the
2d AdS black hole arises as compactification of 3d one \cite{CM}.
The same arguments of Ref. \cite{CM} apply also in the case under
consideration
because
the 4d black hole solution (\ref{h2}) arises as
compactification of the solutions of 5d Einstein-Maxwell gravity
that behave, near the horizon, as AdS$_{3}\times S^{2}$.
Let us consider the 5d Einstein-Maxwell action,
\begin{equation}\label{h9}
A=\int d^5x\ \sqrt{-g}\ \left\{ R- {1\over 4} F^{2}\right\}.
\end{equation}
Compactifying the fifth dimension $x_{4}$, using the ansatz
\begin{eqnarray}\label{l1}
ds^{2}_{(5)}&=&e^{-{4\over \sqrt 3} \hat \phi}dx_{4}^{2}+e^{{2\over \sqrt 3}
\hat \phi}ds^{2}_{(4)},\nonumber\\
F_{\hat \mu, \hat \nu}&=&\left\{ F_{\mu \nu}, F_{4 \nu}\right), F_{4 \nu}=0,
\quad \hat \mu, \hat \nu=0\ldots 4,
\end{eqnarray}
we get the 4d action (\ref{h1}).
The extremal 5d solutions (\ref{l1}) behave near the horizon as
AdS$_{3}\times S^{2}$,
\begin{equation}\label{l2}
ds^{2}_{(5)}={r\over Q}\left( -dt^{2}+ dx_{4}^{2}\right) + \left({Q\over
r}\right)^{2}dr^{2}
+Q^{2} d\Omega^2_{2}.
\end{equation}
Hence, the explanation of the $\sqrt 2$ factor of Ref. \cite{CM} can be
immediately translated to the case under consideration.
Moreover, the expression (\ref{l2}) suggests that the discrepancy
between
thermodynamical and statistical entropy of the 4d, $a=1/\sqrt3$, black
hole could also have a geometrical explanation.
The 3d part of the metric (\ref{l2}) describes a spacetime that is
AdS$_{3}$ with a conical singularity. In fact, if in Eq. (\ref{l2})
$ x_{4}= Q\varphi$, with $0\le\varphi\le 2\pi$, changing coordinates
$r\rightarrow r^{2}/4, \varphi \rightarrow 2 \varphi$, the 3d part of the metric
(\ref{l2}) becomes
\begin{equation}\label{l3}
ds^{2}_{(3)}= - {r^{2}\over 4 Q^{2}} dt^{2}+{4 Q^{2}\over r^{2}}
dr^{2}+r^{2}d \varphi^{2},
\end{equation}
but with $0\le\varphi\le \pi$.
\section {Conclusions}
The dimensional reduction of 4d heterotic string theory presented in
this paper has shown once again that 2d dilaton gravity models can be
used as
a simplified description that retain the relevant information about
the 4d physics. The class of 2d dilaton gravity models
we have derived gives a general description
of excitations near the extremal 4d heterotic black hole.
The geometrical structures, thermodynamical features and the duality
symmetries of the 4d theory become much simpler when translated in
the 2d context.
Particularly interesting are those 2d models that admit $AdS_{2}$ as
solution. In this case one can use the AdS/CFT duality to compute the
statistical entropy of the near-extreme 4d black hole. We have performed this
calculation for the $a=1/\sqrt 3$ black hole but in principle the same
should be possible for the $a=0$ black hole.
The near-horizon geometry factorizes also for $a=0$ as AdS$_{2}
\times S^{2}$ and the excitations near extremality are now described
by the model (\ref{f9}). Differently from the $a=1/\sqrt 3$
case, where we have a linear varying $\exp(-2\Phi)$, for $a=0$ the dilaton is
constant.
A constant dilaton makes a black hole interpretation of the solutions very
difficult,
at least from the 2d point of view. One cannot use the
arguments of Ref. \cite{CM} to compute the statistical entropy of the
black hole.
This difficulties of the $a=0$ case (actually the most
interesting case from the string point of view) are connected
with a peculiarity, mentioned in the introduction, of the $a=0$ case,
namely the existence of a mass gap separating the extremal configuration
from the continuous part of the spectrum. This implies that the
finite-energy excitations near extremality are suppressed
\cite{MMS}.
Probably, this behavior is related with other puzzling features
of the AdS$_{2}$/CFT$_{1}$ correspondence \cite{MMS,st1, CM}.
.
|
\section{Introduction}
\setcounter{equation}{0}
A deeper understanding of flavor mixing and $CP$
violation, observed in the weak interactions,
remains one of the major challenges in particle physics. In the
standard electroweak theory with three quark families
the phenomenon of flavor mixing is described by a $3\times 3$
unitary matrix, which can be expressed in terms of four
independent parameters,
usually taken to be three rotation angles and one complex phase.
There seems no way to obtain any further
information about these parameters within the standard model.
Any attempt to do so would require new physical inputs
which are beyond the standard model.
At the present time it seems hopeless to find a complete
solution to the fermion mass and flavor mixing problem by
theoretical insight alone. One can hope, however, to detect a
specific order in the tower of fermion masses and the four
parameters of quark flavor mixing, especially in observing
links between the parameters of the flavor mixing and the mass
eigenvalues. That such links should exist, seems obvious to us.
Like in any quantum mechanical system the mixing pattern of the
states will influence the pattern of the mass eigenvalues, and
vice versa. One possible way to make these links more
transparent is to look for specific symmetry limits, e.g.,
by setting parameters, which are observed to be small, to
zero and to study the situation in the symmetry limit first.
Following such an approach, we shall demonstrate that (a)
a specific description of quark flavor mixing can be derived,
(b) two of the three flavor mixing angles are related
directly to the quark mass ratios $m_u/m_c$ and $m_d/m_s$,
and (c) the unitarity triangle of quark mixing
related to $CP$ violation in
$B$-meson decays is fixed in terms of these mass ratios and the
modulus of the Cabibbo transition element $|V_{us}|$.
Furthermore we shall give arguments why an inner angle
of the unitarity triangle (angle $\alpha$)
should be equal to $90^{\circ}$ or close to $90^{\circ}$.
The ``standard'' parametrization of the flavor mixing matrix
(advocated by the Particle Data Group \cite{PDG98}) and
the original Kobayashi-Maskawa parametrization \cite{KM}
were introduced
without taking possible links between the quark masses
and the flavor mixing parameters into account. The
parametrization introduced by us some time ago \cite{FX97,FX98}
is based on such a connection, although the
specific relations between flavor mixing angles and
quark masses might be more complicated than
commonly envisaged. It is a parametrization which allows
to interpret the phenomenon of flavor mixing as an
evolutionary or tumbling process. In the limit in which the masses
of the light quarks $(u,d)$ and the medially light
quarks $(c,s)$ are set to zero, while the heavy quarks
$(t,b)$ acquire their masses, there is no flavor mixing
\cite{F87}. Once the masses of the $(c,s)$ quarks are
introduced, while the $(u,d)$ quarks remain massless, the
flavor mixing is reduced to an admixture between two families,
described by one angle $\theta$. As soon as the $u$-
and $d$-quark masses are introduced as small perturbations,
the full flavor mixing matrix involving a complex
phase parameter and two more mixing angles
$(\theta_{\rm u}, \theta_{\rm d})$ appears.
These angles can be interpreted as rotations between
the states $(u,c)$ and $(d,s)$, respectively.
In either the ``standard''
parametrization or the Kobayashi-Maskawa representation,
however, such specific limits are difficult to
consider. For this reason we proceed to describe the flavor
mixing by use of the parametrization given in Ref. \cite{FX97}.
\section{The flavor mixing matrix}
\setcounter{equation}{0}
In the standard model or those extensions which have no
flavor-changing right-handed currents, it is always
possible to choose a basis of flavor space in which
the up- and down-type quark mass matrices are
hermitian. Without loss of any generality the (1,3) and
(3,1) elements of both mass matrices
can further be arranged, through a common unitary
transformation, to be zero \cite{FX97}.
Then one is left
with hermitian quark mass matrices of the form
\begin{equation}
M_{\rm q} \; =\; \left (\matrix{
E_{\rm q} & D_{\rm q} & {\bf 0} \cr
D^*_{\rm q} & C_{\rm q} & B_{\rm q} \cr
{\bf 0} & B^*_{\rm q} & A_{\rm q} \cr}
\right ) \; ,
\end{equation}
where q = u (up) or d (down), and the hierarchy
$|A_{\rm q}| \gg |B_{\rm q}|, |C_{\rm q}|
\gg |D_{\rm q}|, |E_{\rm q}|$ is generally expected.
In this basis, there is no direct mixing between
the heavy $t$ (or $b$) quark and the light $u$ (or $d$)
quark in $M_{\rm u}$ (or $M_{\rm d}$), i.e., the
quark mass matrix is close to the well-known form of
``nearest-neighbour'' interactions \cite{F78}.
A mass matrix of the type (2.1) can in the absence
of complex phases be diagonalized by a $3\times 3$
orthogonal matrix, described only by two rotation
angles in the hierarchy limit of quark masses \cite{F79}.
First, the off-diagonal element $B_{\rm q}$ is rotated away
by a rotation matrix $R_{23}$
between the second and third families.
Then the element $D_{\rm q}$ is rotated away by a
transformation $R_{12}$ between the first and second families.
No rotation between the first and third families is
necessary in either the limit $m_u\rightarrow 0$,
$m_d\rightarrow 0$ or the limit $m_t\rightarrow
\infty$, $m_b\rightarrow \infty$. Lifting such a
hierarchy limit, which is not far from the reality,
one needs an additional transformation
$R_{31}$ with a tiny rotation angle
to fully diagonalize $M_{\rm q}$.
Note, however, that the rotation sequence
$(R^{\rm u}_{12} R^{\rm u}_{23}) (R^{\rm d }_{12}
R^{\rm d }_{23})^{\rm T}$ is enough to describe the $3\times 3$ real
flavor mixing matrix, as the effects of $R^{\rm u}_{31}$
and $R^{\rm d}_{31}$ can always be absorbed into
this sequence through redefining the relevant rotation
angles. By introducing a complex phase angle into
the rotation combination $(R^{\rm u}_{23}) (R^{\rm d}_{23})^{\rm T}$,
we finally arrive at the following representation of
quark flavor mixing \cite{FX97}:
\begin{eqnarray}
V & = & \left ( \matrix{
c_{\rm u} & s_{\rm u} & 0 \cr
-s_{\rm u} & c_{\rm u} & 0 \cr
0 & 0 & 1 \cr } \right ) \left ( \matrix{
e^{-{\rm i}\varphi} & 0 & 0 \cr
0 & c & s \cr
0 & -s & c \cr } \right ) \left ( \matrix{
c_{\rm d} & -s_{\rm d} & 0 \cr
s_{\rm d} & c_{\rm d} & 0 \cr
0 & 0 & 1 \cr } \right ) \nonumber \\ \nonumber \\
& = & \left ( \matrix{
s_{\rm u} s_{\rm d} c + c_{\rm u} c_{\rm d} e^{-{\rm i}\varphi} &
s_{\rm u} c_{\rm d} c - c_{\rm u} s_{\rm d} e^{-{\rm i}\varphi} &
s_{\rm u} s \cr
c_{\rm u} s_{\rm d} c - s_{\rm u} c_{\rm d} e^{-{\rm i}\varphi} &
c_{\rm u} c_{\rm d} c + s_{\rm u} s_{\rm d} e^{-{\rm i}\varphi} &
c_{\rm u} s \cr
- s_{\rm d} s & - c_{\rm d} s & c \cr } \right ) \; ,
\end{eqnarray}
where $s_{\rm u} \equiv \sin\theta_{\rm u}$,
$c_{\rm u} \equiv \cos\theta_{\rm u}$, etc. The three
mixing angles can all be arranged to lie in the first
quadrant, i.e., all $s_{\rm u}$, $s_{\rm d}$, $s$ and
$c_{\rm u}$, $c_{\rm d}$, $c$ are positive.
The phase $\varphi$ may in general take all values
between 0 and $2\pi$. Clearly $CP$ violation is present, if
$\varphi \neq 0$ or $\pi$.
Although we have derived in a heuristic way the particular
description of the flavor mixing matrix (2.2) from the
hierarchical mass matrix (2.1), we should like to emphasize
that (2.2) is a possible way to describe any mixing matrix,
one out of nine inequivalent representations classified
in Ref. \cite{FX98}.
If the phase $\varphi$ in $V$ is disregarded, the resulting rotation matrix
(obtained from (2.2) for $\varphi =0$) is just the one
used originally by Euler; i.e., the angles $\theta$, $\theta_{\rm u}$
and $\theta_{\rm d}$ correspond to the usual Euler angles \cite{EUL}. Note
that this is not the case for other representations of the flavor
mixing matrix given in the literature \cite{KM,Others}.
The representation given in (2.2) can be interpreted as follows.
First, a rotation by the angle $\theta_{\rm d}$ takes place in the
plane defined by the $d$ and $s$ quarks. It is followed by
a rotation (angle $\theta$) in the $b$--$s'$ plane, where $s'$ denotes the
superposition $s' = d \sin\theta_{\rm d} + s \cos\theta_{\rm d}$. At
the same time the orthogonal
state $d' = d \cos\theta_{\rm d} - s \sin\theta_{\rm d}$
is multiplied by the phase factor $e^{-{\rm i}\varphi}$. Finally a rotation
(angle $\theta_{\rm u}$) is applied in the
$1$--$2$ plane (about the new third axis).
The sequence of rotations corresponds just to the Euler sequence \cite{EUL}:
$R_{12} R_{23} R^{\rm T}_{12}$.
On the other hand, the original Kobayashi-Maskawa
representation \cite{KM} corresponds to the sequence
$R_{23} R_{12} R^{\rm T}_{23}$, while the ``standard''
representation \cite{PDG98}
corresponds to the sequence $R_{23} R_{31} R_{12}$ (see also
the classifications given in Ref. \cite{FX98}). Although all
descriptions of the flavor mixing matrix are mathematically
equivalent, we emphasize that the Euler
sequence $R_{12} R_{23} R^{\rm T}_{12}$ is physically of
particular interest, as it involves the rotation matrices $R_{12}$
and $R^{\rm T}_{12}$, which describe the rotations in the light
quark sector, in a symmetric way.
Since the flavor mixing matrix
acts between the quark mass eigenstates ${\cal U} = (u, c, t)$ and
${\cal D} = (d, s, b)$, one could absorb the two $R_{12}$ rotations
in a redefinition of the quark fields.
The charged weak transition term can be rewritten as follows:
\begin{equation}
\overline{\cal U}_{\rm L} ~ V ~ {\cal D}_{\rm L} \; =\;
\overline{(u, ~ c, ~ t)}^{~}_{\rm L} ~ V ~ \left ( \matrix{
d \cr s \cr b \cr} \right )_{\rm L} \; =\;
\overline{(u', ~ c', ~ t)}^{~}_{\rm L} \left ( \matrix{
e^{-{\rm i}\varphi} & 0 & 0 \cr
0 & c & s \cr
0 & -s & c \cr} \right )
\left ( \matrix{
d' \cr s' \cr b \cr} \right )_{\rm L} \; ,
\end{equation}
where $u' = u \cos\theta_{\rm u} - c \sin\theta_{\rm u}$ and
$c' = c \cos\theta_{\rm u} + u \sin\theta_{\rm u}$.
Thus the angles $\theta_{\rm u}$ and $\theta_{\rm d}$ describe the
corresponding rotations in the $(u, c)$ and $(d, s)$ systems.
We should like to emphasize that the angles $\theta_{\rm u}$ and
$\theta_{\rm d}$ can directly be measured from weak decays of
$B$ mesons and from $B^0$-$\bar{B}^0$ mixing.
An analysis of the present experimental
data yields \cite{PRS98}: $\theta_{\rm u} = 4.87^{\circ} \pm 0.86^{\circ}$
and $\theta_{\rm d} = 11.71^{\circ} \pm 1.09^{\circ}$. Taking the
central values for illustration, one has
\begin{eqnarray}
d' & = & d \cos\theta_{\rm d} ~ - ~ s \sin\theta_{\rm d} \; \approx \;
0.979 d ~ - ~ 0.203 s \; , \nonumber \\
s' & = & d \sin\theta_{\rm d} ~ + ~ s \cos\theta_{\rm d} \; \approx \;
0.203 d ~ + ~ 0.979 s \; ; \nonumber \\
u' & = & u \cos\theta_{\rm u} ~ - ~ c \sin\theta_{\rm u} \; \approx \;
0.996 u ~ - ~ 0.085 c \; , \nonumber \\
c' & = & u \sin\theta_{\rm u} ~ + ~ c \cos\theta_{\rm u} \; \approx \;
0.085 u ~ + ~ 0.996 c \; .
\end{eqnarray}
The question, about whether these mixtures of mass eigenstates have a
specific physical meaning, arises. This will be discussed in some more detail
below. Due to the symmetric structure of our mixing matrix (2.2),
we are able to interpret the $\theta_{\rm d}$ and $\theta_{\rm u}$
rotations as specific transformations of the corresponding
mass eigenstates. Such an interpretation is
not possible for the third rotation given by $\theta$, measured to
be $2.30^{\circ} \pm 0.09^{\circ}$ \cite{PRS98}. This rotation takes
place between the third family of the massive quarks and the $c'$
and $s'$ states.
One interpretation would be to associate the rotation of $\theta$
with a transformation among $b$ and $s'$. Another possibility is to
describe the effect as a rotation among $t$ and $c'$. However, one
could also write $\theta$ as a difference of two other angles, and describe
the mixing effect as a combination of a rotation in the $(b, s')$
system and a rotation in the $(t, c')$ system. Thus a unique
interpretation
does not exist. We remark that the asymmetry between the $\theta$
rotation on the one hand and the $\theta_{\rm u}$ and $\theta_{\rm d}$
rotations on the other hand is a direct consequence of our flavor
mixing matrix (which is in turn related to the hierarchical structure
of the mass spectrum) and is linked to the fact that there exist
three different quark families.
It is worthwhile to point out
the similarity and difference between
our new parametrization and the Kobayashi-Maskawa parametrization, which
both result from rotations in the 1--2 and 2--3 planes
(i.e., $R_{12}$ and $R_{23}$), in the description of quark
flavor mixing.
To make a comparison, we write out the latter as follows:
\begin{eqnarray}
V_{\rm KM} & = &
\left (\matrix{
1 & 0 & 0 \cr
0 & c_2 & s_2 \cr
0 & -s_2 & c_2 \cr} \right )
\left (\matrix{
c_1 & s_1 & 0 \cr
-s_1 & c_1 & 0 \cr
0 & 0 & e^{-{\rm i}\delta} \cr} \right )
\left (\matrix{
1 & 0 & 0 \cr
0 & c_3 & -s_3 \cr
0 & s_3 & c_3 \cr} \right ) \; \nonumber \\ \nonumber \\
& = & \left ( \matrix{
c_1 & s_1 c_3 & -s_1 s_3 \cr
-s_1 c_2 & c_1 c_2 c_3 + s_2 s_3 e^{-{\rm i} \delta} &
-c_1 c_2 s_3 + s_2 c_3 e^{-{\rm i}\delta } \cr
s_1 s_2 & -c_1 s_2 c_3 + c_2 s_3 e^{-{\rm i}\delta} &
c_1 s_2 s_3 + c_2 c_3 e^{-{\rm i}\delta} \cr } \right ) \; ,
\end{eqnarray}
where $s_1 \equiv \sin\theta_1$, $c_1 \equiv \cos\theta_1$, etc.
The mixing angles $(\theta_{\rm u}, \theta_{\rm d}, \theta)$ are related
to $(\theta_1, \theta_2,
\theta_3)$ simply through the product of $|V_{ub}|$ and
$|V_{td}|$, i.e.,
$s_{\rm u} s_{\rm d} s^2 = s^2_1 s_2 s_3$ holds. One can also link
the phase parameter $\varphi$ to $\delta$ with the help of the
common rephasing-invariant measure of $CP$ violation \cite{J85};
i.e.,
\begin{eqnarray}
{\cal J} & = & s_{\rm u} c_{\rm u} s_{\rm d} c_{\rm d}
s^2 c \sin\varphi \nonumber \\
& = & s^2_1 c_1 s_2 c_2 s_3 c_3 \sin\delta \; .
\end{eqnarray}
With no fine-tuning of the relevant mixing angles, we arrive at the
equality between $\varphi$ and $\delta$ to an excellent degree of
accuracy:
\begin{equation}
\frac{\sin \varphi}{\sin \delta} \; =\; \frac{c_1 ~ c_2 ~ c_3}{c_{\rm u}
~ c_{\rm d} ~ c} \; =\; 1 ~ - ~ O(\lambda^2) \; ,
\end{equation}
where $\lambda \approx s_{\rm d} \approx s_1 \approx 0.2$.
Therefore a large $CP$-violating phase (close to $90^{\circ}$),
as required either
phenomenologically \cite{FX97} or in a specific dynamical scheme
\cite{FX95,Weyers}, must manifest itself in both (2.2) and (2.5).
The difference between these two representations is however
significant. For example, the Kobayashi-Maskawa parametrization
starts from the second
largest matrix element of $V$ (i.e., $|V_{ud}|$ instead of $|V_{tb}|$)
and leads to quite complicated results
for the ratios $|V_{ub}/V_{cb}|$ and $|V_{td}/V_{ts}|$.
As summarized in Refs. \cite{FX97,FX98}, the
new parametrization (2.2) has a number of advantages over all
the others in the
study of heavy flavor decays and quark mass matrices.
Its usefulness will be seen more clearly in the present work.
As an example we explore the interesting connection between
our parametrization (2.2) and the unitarity triangle of quark
mixing defined by the orthogonality relation
\begin{equation}
V^*_{ub}V_{ud} ~ + ~ V^*_{cb}V_{cd} ~ + ~ V^*_{tb}V_{td} \; =\; 0
\end{equation}
in the complex plane. The inner angles of this triangle,
usually denoted as
\begin{eqnarray}
\alpha & = & \arg \left ( - \frac{V^*_{tb}V_{td}}{V^*_{ub}V_{ud}}
\right ) \; , \nonumber \\
\beta & = & \arg \left (- \frac{V^*_{cb}V_{cd}}{V^*_{tb}V_{td}}
\right ) \; , \nonumber \\
\gamma & = & \arg \left (- \frac{V^*_{ub}V_{ud}}{V^*_{cb}V_{cd}}
\right ) \; ,
\end{eqnarray}
can be determined from some $CP$-violating asymmetries in
$B$-meson decays \cite{BB}. Current data indicate that
the unitarity triangle (2.8) is congruent, to a good degree of
accuracy, with another unitarity triangle defined
by the orthogonality relation
$V^*_{td}V_{ud} + V^*_{ts}V_{us} + V^*_{tb}V_{ub} =0$
in the complex plane \cite{Xing96}. In view of
the approximate congruency between two unitarity triangles and
the smallness of three mixing angles, we find
that the parametrization (2.2) takes an instructive
leading-order form:
\begin{equation}
V \; \approx \; \left (\matrix{
e^{-{\rm i}\alpha} & s^{~}_{\rm C} e^{{\rm i}\gamma} & s_{\rm u} s \cr
s^{~}_{\rm C} e^{{\rm i}\beta} & 1 & s \cr
-s_{\rm d} s & -s & 1 \cr} \right ) \; ,
\end{equation}
where $s^{~}_{\rm C} \equiv \sin \theta_{\rm C} \approx
|s_{\rm u} - s_{\rm d} e^{-{\rm i}\varphi}|$ with $\theta_{\rm C}$
denoting the Cabibbo rotation angle \cite{Cabibbo}.
Clearly $\alpha \approx \varphi$ holds as a straightforward result of
(2.10). In this approximation $|V^*_{ub}V_{ud}|$, $|V^*_{cb}V_{cd}|$
and $|V^*_{tb}V_{td}|$, three sides of the unitarity triangle (2.8),
are rescaled to $s_{\rm u}$, $s_{\rm d}$ and $s^{~}_{\rm C}$
respectively. The latter are three sides of a new triangle with
smaller area ($\approx s_{\rm u}s_{\rm d} \sin\alpha/2$),
which will subsequently be referred to as the ``light-quark triangle''
in the heavy quark limit ($m_t\rightarrow \infty$,
$m_b\rightarrow \infty$).
The values of $\alpha$, $\beta$ and $\gamma$ can
therefore be given in terms of $s_{\rm u}$, $s_{\rm d}$ and
$s^{~}_{\rm C}$ with the help of the cosine theorem. In
particular, relations like \cite{FX97}
\begin{equation}
\sin\alpha ~ : ~ \sin\beta ~ : ~ \sin\gamma \; \approx \;
s^{~}_{\rm C} ~ : ~ s_{\rm u} ~ : ~ s_{\rm d} \;
\end{equation}
may directly be confronted with the upcoming data on $CP$
asymmetries in $B$ decays \cite{CDF}. Motivated by these
interesting results, we shall investigate the role that the
light quark sector plays in $CP$ violation for a variety of
realistic textures of quark mass matrices.
\section{Symmetry limits of quark masses}
\setcounter{equation}{0}
Going farther from the previous discussions \cite{FX97,FX98},
we remark two useful limits of quark masses
and analyze their corresponding consequences on flavor mixing.
\subsection{The chiral limit of quark masses}
In the limit $m_u \rightarrow 0$, $m_d \rightarrow 0$ (``chiral
limit''), where both the up and down quark mass
matrices have zeros in
the positions $(1,1)$, $(1,2)$, $(2,1)$, $(1,3)$ and $(3,1)$ (see also
Ref. \cite{F87}), the flavor mixing angles
$\theta_{\rm u}$ and $\theta_{\rm d}$ vanish. Only the $\theta$ rotation
affecting the heavy quark sector remains, i.e.,
the flavor mixing matrix effectively takes the form
\begin{equation}
\hat{V} \; =\; \left ( \matrix{
\cos\hat{\theta} & \sin\hat{\theta} \cr
- \sin\hat{\theta} & \cos\hat{\theta} \cr } \right ) \; ,
\end{equation}
where $\hat{\theta}$ denotes the value of $\theta$ which one
obtains in
the limit $\theta_{\rm u} \rightarrow 0$, $\theta_{\rm d} \rightarrow
0$. We see that $\hat{V}$ is a real orthogonal matrix, arising naturally
from $V$ in the chiral limit.
The flavor mixing angle $\hat{\theta}$ can be derived from hermitian
quark mass matrices of the following general form (in the limit $m_u
\rightarrow 0$, $m_d \rightarrow 0$):
\begin{equation}
\hat{M}_{\rm q} \; =\; \left ( \matrix{
\hat{C}_{\rm q} & \hat{B}_{\rm q} \cr
\hat{B}^*_{\rm q} & \hat{A}_{\rm q} \cr } \right ) \; ,
\end{equation}
where $|\hat{A}_{\rm q}| \gg |\hat{B}_{\rm q}| , |\hat{C}_{\rm q}|$;
and q = u (up) or d
(down). Note that the phase difference between $\hat{B}_{\rm u}$ and
$\hat{B}_{\rm d}$, denoted as $\kappa \equiv \arg (\hat{B}_{\rm u})
- \arg (\hat{B}_{\rm d})$, has no effect on $CP$ symmetry
in the chiral limit, but it may affect the magnitude of
$\hat{\theta}$. It is known that current data on the top-quark mass
and the $B$-meson lifetime disfavor the special case $\hat{C}_{\rm u}
= \hat{C}_{\rm d} =0$ for $\hat{M}_{\rm u}$ and
$\hat{M}_{\rm d}$ (see, e.g., Ref. \cite{FX95}),
hence we take $\hat{C}_{\rm q} \neq 0$ and define a ratio $\hat{r}_{\rm q}
\equiv |\hat{B}_{\rm q}|/\hat{C}_{\rm q}$ for convenience. Then we
can obtain the flavor mixing angle $\hat{\theta}$, in terms of the
quark mass ratios $m_c/m_t$, $m_s/m_b$ and the parameters
$\hat{r}_{\rm u}$, $\hat{r}_{\rm d}$, by diagonalizing the mass
matrices in (3.2). In the next-to-leading order approximation,
$\sin\hat{\theta}$ reads
\begin{equation}
\sin\hat{\theta} \; = \; \left | \hat{r}_{\rm d} \frac{m_s}{m_b} \left (1 -
\hat{\delta}_{\rm d} \right ) ~ - ~ \hat{r}_{\rm u} \frac{m_c}{m_t} \left (1 -
\hat{\delta}_{\rm u} \right ) e^{{\rm i} \kappa} \right | \; ,
\end{equation}
where two correction terms are given by
\begin{eqnarray}
\hat{\delta}_{\rm u} & = & \left (1 + \hat{r}^2_{\rm u}
\right ) \frac{m_c}{m_t} \; , \nonumber \\
\hat{\delta}_{\rm d} & = & \left (1 + \hat{r}^2_{\rm d} \right )
\frac{m_s}{m_b}
\; .
\end{eqnarray}
In view of the fact $m_s/m_b \sim O(10) ~ m_c/m_t$ from current data
\cite{PDG98,Leut}, we find that the
flavor mixing angle $\hat{\theta}$ is primarily linked to $m_s/m_b$
provided $|\hat{r}_{\rm u}| \approx |\hat{r}_{\rm d}|$. Note that in specific
models, e.g., those describing the mixing between the second and third
families as an effect related to the breaking of an underlying ``democratic
symmetry'' \cite{Democracy,New},
the ratios $\hat{r}_{\rm u}$ and $\hat{r}_{\rm d}$
are purely algebraic numbers (such as
$|\hat{r}_{\rm u}| = |\hat{r}_{\rm d}| = 1/\sqrt{2}$
or $\sqrt{2}$).
For illustration, we take $\hat{r}_{\rm u} = \hat{r}_{\rm d} \equiv
\hat{r}$ to fit the experimental result $\sin\hat{\theta} = 0.040 \pm 0.002$
with the typical inputs $m_b/m_s = 26 - 36$ and $m_t/m_c \sim 250$.
It is found that the favored value of $|\hat{r}|$ varies in the range
1.0 -- 2.5, dependent weakly on the phase parameter $\kappa$.
Note that both $m_s/m_b$ and $m_c/m_t$ evolve with the
energy scale (e.g., from the weak scale $\mu \sim 10^2$ GeV
to a superhigh scale $\mu \sim 10^{16}$ GeV,
or vice versa), therefore $\tilde{\theta}$ itself is also
scale-dependent.
\subsection{The heavy quark limit}
The limit $m_t \rightarrow \infty$, $m_b \rightarrow \infty$ is
subsequently referred to as the ``heavy quark limit''. In this limit,
in which the $(3,3)$ elements of the up and down mass matrices formally
approach infinity but
all other matrix elements are fixed, the angle $\theta$ vanishes.
The flavor mixing matrix,
which is nontrivial only in the light quark sector, takes the form:
\begin{eqnarray}
\tilde{V} & = & \left ( \matrix{
\tilde{c}_{\rm u} & \tilde{s}_{\rm u} \cr
-\tilde{s}_{\rm u} & \tilde{c}_{\rm u} \cr } \right )
\left ( \matrix{
e^{-{\rm i}\tilde{\varphi}} & 0 \cr
0 & 1 \cr } \right ) \left ( \matrix{
\tilde{c}_{\rm d} & -\tilde{s}_{\rm d} \cr
\tilde{s}_{\rm d} & \tilde{c}_{\rm d} \cr } \right )
\nonumber \\ \nonumber \\
& = & \left ( \matrix{
\tilde{s}_{\rm u} \tilde{s}_{\rm d} + \tilde{c}_{\rm u} \tilde{c}_{\rm d}
e^{-{\rm i}\tilde{\varphi}} &
\tilde{s}_{\rm u} \tilde{c}_{\rm d} - \tilde{c}_{\rm u} \tilde{s}_{\rm d}
e^{-{\rm i}\tilde{\varphi}} \cr
\tilde{c}_{\rm u} \tilde{s}_{\rm d} - \tilde{s}_{\rm u} \tilde{c}_{\rm d}
e^{-{\rm i}\tilde{\varphi}} &
\tilde{c}_{\rm u} \tilde{c}_{\rm d} + \tilde{s}_{\rm u} \tilde{s}_{\rm
d} e^{-{\rm i}\tilde{\varphi}} \cr } \right ) \; .
\end{eqnarray}
where $\tilde{s}_{\rm u} = {\rm sin} \tilde{\theta }_{\rm u},
\tilde{c}_{\rm u} = {\rm cos} \tilde{\theta}_{\rm u}$, etc. The angles
$\tilde{\theta}_{\rm u}$ and $\tilde{\theta}_{\rm d}$ are the
values for $\theta_{\rm u}$ and $\theta_{\rm d}$
obtained in the heavy quark limit. Since the $(t, b)$ system is decoupled from
the $(c, s)$ and $(u, d)$ systems, the flavor mixing can be described as in
the case of two families. Therefore the mixing matrix $\tilde{V}$ is
effectively given in terms of only a single rotation angle,
the Cabbibo angle $\theta_{\rm C}$ \cite{Cabibbo}:
\begin{equation}
\sin \theta_{\rm C} = \mid \tilde{s}_{\rm u} \tilde{c}_{\rm d}
~ - ~ \tilde{c}_{\rm u} \tilde{s}_{\rm d} ~ e^{-{\rm i}
\tilde{\varphi}} \mid \; .
\end{equation}
Of course $\tilde{V}(\theta_{\rm C})$
is essentially a real matrix, because its complex
phases can always be rotated away by redefining the quark fields.
We should like to stress that the heavy quark limit, which carries
the flavor mixing matrix $V$ to its simplified form $\tilde{V}$, is not far
from the reality, since $1 - c \approx 0.1 \%$ holds \cite{FX97}.
Therefore $\theta_{\rm u}$,
$\theta_{\rm d}$ and $\varphi $ are expected to
approach $\tilde{\theta}_{\rm u}$, $\tilde{\theta}_{\rm d}$ and
$\tilde{\varphi}$ rapidly,
as $\theta \rightarrow 0$, corresponding to
$m_t \rightarrow \infty $ and $m_b \rightarrow \infty$. However, the concrete
limiting behavior depends on the specific algebraic
structure of the up and down mass matrices.
If two hermitian mass matrices have the parallel hierarchy with texture zeros
in the (1,1) (2,2), (1,3) and (3,1) elements, for example, the magnitude of
$\theta $ is suppressed by the terms proportional to $m_t^{-1/2}$ and
$m_b^{-1/2}$ \cite{F79}; and
if the (2,2) elements are kept nonvanishing and comparable in magnitude
with the (2,3) and (3,2) elements, then
$\theta $ is dependent
on $m_t^{-1}$ and $m_b^{-1}$ \cite{Democracy,New}.
The angles $\tilde{\theta}_{\rm u}$ and
$\tilde{\theta}_{\rm d}$ as well as the phase
$\tilde{\varphi }$ are well-defined quantities in the heavy quark limit. The
physical meaning of these quantities can be seen more clearly, if we take
into account a specific and realistic model for the Cabibbo-type mixing in
the light quark sector. It is well known that in the absence of the
$u$-quark mass a relation between the Cabibbo angle $\theta_{\rm C}$
and the mass
ration $m_d/m_s$ follows, if the quark mass matrices have the structure:
\begin{eqnarray}
\tilde{M}_{\rm u} & = & \left( \matrix{
~ {\bf 0} & ~ {\bf 0} ~ \cr
~ {\bf 0} & ~ m_c \cr } \right) \; , \nonumber \\
\tilde{M}_{\rm d} & = & \left( \matrix{
{\bf 0} & \tilde{B}_{\rm d} \cr
\tilde{B}^*_{\rm d} & \tilde{A}_{\rm d} \cr } \right) \; .
\end{eqnarray}
The diagonalization of $\tilde{M}_{\rm d}$ leads to the relation
${\rm tan } \theta_{\rm C} = \sqrt{m_d/m_s}$ . The texture-zero pattern
of $\tilde{M}_{\rm d}$,
i.e., the vanishing of its (1,1) element, is already
present in certain classes of
models (see, e.g., Refs. \cite{F77,Weinb}).
The relation for the Cabibbo angle is known to
agree very well with the experimental observation. For numerical discussions,
we make use of the quark masses $m_u = (5.1 \pm 0.9)$ MeV, $m_d = (9.3 \pm
1.4)$ MeV, $m_s = (175 \pm 25)$ MeV and $m_c = (1.35 \pm 0.05)$ GeV at the
scale $\mu =1$ GeV \cite{Leut}. Then one finds
$\theta_{\rm C} = 13.0^{\circ } \pm 1.8^{\circ}$ or
$\sin \theta_{\rm C} = 0.225 \pm 0.031$, consistent with the observed
value of $|V_{us}|$ (i.e., $0.217 \leq |V_{us}| \leq 0.224$
\cite{PDG98}).
The situation will change once $m_u$ is introduced, i.e.,
$\tilde{M}_{\rm u}$ takes the same
form as $\tilde{M}_{\rm d}$ given in (3.7). In this case
the mass matrices result in the
following relation \cite{F79}:
\begin{equation}
\sin \theta_{\rm C} \; = \; \mid R_{\rm u} ~ - ~ R_{\rm d} ~ e^{-{\rm i}
\psi} \mid \; ,
\end{equation}
where
\begin{eqnarray}
R_{\rm u} & = & \sqrt{\frac{m_u}{m_u + m_c}} \, \sqrt{\frac{m_s}{m_d + m_s}}
\;\; , \nonumber \\
R_{\rm d} & = & \sqrt{\frac{m_c}{m_u + m_c}} \, \sqrt{\frac{m_d}{m_d +
m_s}} \; \; ,
\end{eqnarray}
and $\psi \equiv \arg (\tilde{B}_{\rm u}) - \arg (\tilde{B}_{\rm d})$
denotes the relative phase between the off-diagonal elements
$\tilde{B}_{\rm u}$ and $\tilde{B}_{\rm d}$
(in the limit $m_u \rightarrow 0$ this phase can be absorbed through a
redifinition of the quark fields). We find that the same structure for the
Cabibbo-type mixing matrix has been
obtained as in the decoupling limit discussed
above. If we set
\begin{eqnarray}
\tan \tilde{\theta}_{\rm u} & = & \sqrt{\frac{m_u}{m_c}} \; , \nonumber \\
\tan \tilde{\theta}_{\rm d} & = & \sqrt{\frac{m_d}{m_s}} \; ,
\end{eqnarray}
and $\tilde{\varphi} = \psi$ for (3.6), then the result in (3.8)
and (3.9) can exactly be reproduced.
\begin{figure}[t]
\begin{picture}(400,160)(-90,210)
\put(80,300){\line(1,0){150}}
\put(80,300.5){\line(1,0){150}}
\put(150,285.5){\makebox(0,0){$\sin\theta_{\rm C}$}}
\put(80,300){\line(1,3){21.5}}
\put(80,300.5){\line(1,3){21.5}}
\put(80,299.5){\line(1,3){21.5}}
\put(71,333){\makebox(0,0){$R_{\rm u}$}}
\put(230,300){\line(-2,1){128}}
\put(230,300.5){\line(-2,1){128}}
\put(178,343.5){\makebox(0,0){$R_{\rm d}$}}
\put(107.5,348.5){\makebox(0,0){$\tilde{\varphi}$}}
\end{picture}
\vspace{-2.6cm}
\caption{The light-quark triangle (LT) in the complex plane.}
\end{figure}
Indeed the relation in (3.6) or (3.8) defines a triangle in the complex plane
(see Fig. 1 for illustration), which will be denoted as the ``light-quark
triangle''(LT). Taking into account the central values of the
Cabibbo angle ($\sin \theta_{\rm C} = |V_{us}| = 0.2205$) and the light
quark mass ratios $\left( m_s / m_d = 18.8 \right. $ and
$\left. m_c /m_u = 265 \right)$, we
can calculate the phase parameter from (3.8) and obtain
$\tilde{\varphi} = \psi
\approx 79^{\circ}$. If we allow the mass ratios
and $\theta_{\rm C}$ to vary in
their ranges given above, then $\tilde{\varphi}$ may vary in the range
$38^{\circ} - 115^{\circ}$. We find that $\tilde{\varphi}$ has a good
chance to be around $90^{\circ}$ (see also Ref. \cite{FX95}).
The case $\tilde{\varphi} \approx 90^{\circ}$ (i.e., the LT is
rectangular) is of special interest, as we shall see later, since it
implies that the area of the unitarity triangle of flavor mixing
takes its maximum value for the fixed quark mass ratios -- in this
sense, the $CP$ symmetry of weak interactions would be maximally violated.
It is worth remarking that the quark mass ratios
$m_d/m_s$ and $m_u/m_c$ are essentially
independent of the renormalization-group effect from the weak scale
to a superhigh scale (or vice versa), so is the Cabibbo angle
$\theta_{\rm C}$. As a result the sides and angles of the LT
are to a very good degree of accuracy
scale-independent. This interesting feature of the light
quark sector implies
that the prediction for $\tilde{\theta}_{\rm u}$
and $\tilde{\theta}_{\rm d}$ from quark mass matrices at any
high scale (e.g., $\mu \sim 10^{16}$ GeV) can directly
be confronted with the low-scale experimental data.
The two symmetry limits discussed above are both not far from
the reality, in which the strong hierarchy of quark masses
($m_u\ll m_c \ll m_t$ and $m_d\ll m_s\ll m_b$) has been
observed. They will serve as a guide in the subsequent
discussions about generic quark mass matrices and their
consequences on flavor mixing.
\section{Analysis of generic mass matrices}
\setcounter{equation}{0}
Now we return to the case of three quark families.
In the standard model or its extensions which have no flavor-changing
right-handed currents,
one can always adopt a basis of flavor space in which both the up-
and down-type quark mass matrices are hermitian and have
vanishing (1,3) and (3,1) elements, as shown in (2.1).
Such a basis is of special interest in case of a strong mass hierarchy (as
realized by nature), since no explicit mixing between the very massive
$t$ (or $b$) quark and the very light $u$ (or $d$) quark is introduced.
The mixing can then be regarded as
of the ``nearest neighbour'' type \cite{F78}.
Thus without loss of generality one may discuss the model-independent
properties of flavor mixing and $CP$ violation on the basis of the mass
matrices (2.1), i.e.,
\begin{eqnarray}
M_{\rm u} & = & \left ( \matrix{
E_{\rm u} & D_{\rm u} & {\bf 0} \cr
D^*_{\rm u} & C_{\rm u} & B_{\rm u} \cr
{\bf 0} & B^*_{\rm u} & A_{\rm u} \cr} \right ) \; ,
\nonumber \\
M_{\rm d} & = & \left ( \matrix{
E_{\rm d} & D_{\rm d} & {\bf 0} \cr
D^*_{\rm d} & C_{\rm d} & B_{\rm d} \cr
{\bf 0} & B^*_{\rm d} & A_{\rm d} \cr} \right ) \; .
\end{eqnarray}
The phases of $D_{\rm u,d}$ and $B_{\rm u,d}$ elements are denoted
as $\phi_{D_{\rm u,d}}$ and $\phi_{B_{\rm u,d}}$, respectively.
The phase differences
\begin{eqnarray}
\phi_1 & = & \phi_{D_{\rm u}} - \phi_{D_{\rm d}} \; ,
\nonumber \\
\phi_2 & = & \phi_{B_{\rm u}} - \phi_{B_{\rm d}} \;
\end{eqnarray}
are the source of
$CP$ violation in weak interactions of quarks.
It is clear that $M_{\rm u}$ and $M_{\rm d}$ consist totally of
twelve parameters.
If the hermiticity is not imposed on the arbitrary up and down
mass matrices in the standard model, then they can be taken as the full
``nearest-neighbor'' mixing form with texture zeros in
the (1,1), (2,2), (1,3) and (3,1) positions \cite{Branco}:
\begin{equation}
{\cal M}_{\rm q} \; =\; \left (\matrix{
{\bf 0} & {\cal X}_{\rm q} & {\bf 0} \cr
{\cal X}'_{\rm q} & {\bf 0} & {\cal Y}_{\rm q} \cr
{\bf 0} & {\cal Y}'_{\rm q} & {\cal Z}_{\rm q} \cr } \right ) \; ,
\end{equation}
where $\arg ({\cal X}'_{\rm q}) = \arg ({\cal X}^*_{\rm q})$ and
$\arg ({\cal Y}'_{\rm q}) = \arg ({\cal Y}^*_{\rm q})$ (for q = u and d).
In this special basis the light quarks are assumed to acquire
masses through an interaction with their nearest neighbors.
It is straightforward to find that the non-hermitian
mass matrices ${\cal M}_{\rm u,d}$
have the same number of free parameters as the hermitian
mass matrices $M_{\rm u,d}$, therefore one could be
transformed to the other
\footnote{If one imposes the hermiticity on (4.3) or the
nearest-neighbor mixing on (4.1), then the resultant mass
matrices take the particularly simple form which was first
proposed and discussed
by one of the authors about twenty years ago \cite{F77}.}.
In our point of view the hermitian basis (4.1) is more natural and
will be adopted in the subsequent discussions.
\subsection{Conditions for $CP$ violation}
We first discuss
the necessary and sufficient conditions for $CP$ violation in the standard
electroweak model and clarify some ambiguity associated with this
problem in the literature. As the flavor mixing matrix $V$
is obtained from the diagonalization of the mass matrices
$M_{\rm u}$ and $M_{\rm d}$, there must be some kind of relation
between the parameters of $V$ and $M_{\rm u,d}$. The conditions for
$CP$ violation can be counted either at the level of quark
mass matrices or at the level of the flavor mixing matrix.
One must distinguish between these two different levels.
At the level of quark mass matrices it is obvious that $CP$
symmetry will be violated, if and only if there is at least
one nontrivial phase difference between $M_{\rm u}$ and
$M_{\rm d}$. In other words, ${\rm Im} (M_{{\rm u}ij}
M^*_{{\rm d}ij}) \neq 0$ (for $i,j=1,2,3$ and $i<j$) is the
necessary and sufficient condition for $CP$ violation in the
standard model. If one defines a commutator for $M_{\rm u}$
and $M_{\rm d}$, $[M_{\rm u}, M_{\rm d} ] \equiv {\rm i}
~ {\cal C}$, then it is easy to find
\begin{equation}
\frac{{\cal C}_{ii}}{2} \; =\; {\rm Im} \left ( M_{{\rm u}ij} M^*_{{\rm d} ij}
\right ) ~ + ~ {\rm Im} \left (M_{{\rm u}ik} M^*_{{\rm d} ik}
\right ) \; ,
\end{equation}
for $i,j,k=1,2,3$ but $i\neq j\neq k$. Clearly ${\cal C}_{ii} \neq
0$, if $CP$ symmetry is not conserved.
Note that $CP$ symmetry would be conserved,
if two quarks with the same charge were degenerate in mass
eigenvalues. This is well known, but we shall give a proof here.
We assume the $i$ and $j$ quarks
in the up sector to be degenerate, then they would not be
distinguished from each other by any quantum number. Hence
any linear combination of the mass eigenstates $|i\rangle$ and
$|j\rangle$, e.g., $|i'\rangle$ or $|j'\rangle$ in the form
\begin{equation}
\left ( \matrix{
i' \cr
j' \cr } \right ) \; =\;
\left ( \matrix{
\cos\vartheta ~ e^{+{\rm i} \xi}
& +\sin \vartheta ~ e^{+{\rm i} \zeta} \cr
- \sin \vartheta ~ e^{-{\rm i} \zeta}
& \cos \vartheta ~ e^{-{\rm i} \xi} \cr } \right )
\left ( \matrix{ i \cr j \cr} \right ) \; ,
\end{equation}
remains a mass eigenstate. Without loss of any physical content,
the elements of $M_{\rm u}$ in the $i$-th and $j$-th lines and
rows can be
rearranged by three arbitrary (real) parameters $\vartheta$, $\xi$ and
$\zeta$. This, together with other known freedoms, allows
one to remove all possible phase differences between
$M_{\rm u}$ and $M_{\rm d}$, i.e., ${\rm Im}
\left ( M_{{\rm u} ij} M^*_{{\rm d} ij} \right ) = 0$
(for $i\neq j$) appears.
A similar proof is valid for the down quark sector. We then
conclude that a non-degeneracy between the quarks with the
same charge is a necessary (but not sufficient) condition
for $CP$ violation in the standard model. This condition can be
more explicitly written, in terms of the determinant of $\cal C$,
as follows \cite{J89}:
\begin{equation}
{\rm Det} ~ {\cal C} \; =\; -2 {\cal J} \prod_{i < j}
\left (\lambda_i - \lambda_j \right ) \prod_{\alpha < \beta}
\left (\lambda_\alpha - \lambda_\beta \right ) \; ,
\end{equation}
where $\cal J$ can be found in (2.6),
$\lambda_i$ and $\lambda_\alpha$ denote
the quark mass eigenvalues, and the subscripts $(i,j)$ and
$(\alpha, \beta)$ run over $(u,c,t)$ and $(d,s,b)$ respectively.
However, it should be noted that the parameter $\cal J$ itself
{\it does} depend on the product of two mass-eigenvalue differences
$(\lambda_i - \lambda_j)$ and $(\lambda_\alpha -
\lambda_\beta)$, as we shall
prove in the next subsection.
In this sense $\cal J$ and ${\rm Det} ~ {\cal C}$
contain the same information about $CP$ violation; i.e., the
latter is not more fundamental than the former, contrary to
popular belief.
Now we discuss the condition for $CP$ violation at the level
of the flavor mixing matrix. Of course $CP$ symmetry is
violated, if $V$ contains a nontrivial complex phase which
cannot be removed through the redefinition of quark-field
phases. The most appropriate measure of $CP$ violation (due
to the unitarity of $V$) is the rephasing-invariant parameter
$\cal J$, whose relation with three mixing
angles and the $CP$-violating phase has been given in
(2.6).
Obviously $\cal J$ vanishes if $\varphi =0$ or $\pi$. Note
that for $\theta_{\rm u} =0$ or $\pi/2$ the phase parameter
$\varphi$ can be removed from $V$. Therefore the resultant flavor
mixing matrix is a real $3\times 3$ matrix described by only
two rotation angles ($\theta_{\rm d}$ and $\theta$). A similar
situation will appear if $\theta_{\rm d} =0$, $\pi/2$ or
$\theta =0$, $\pi/2$. The necessary and sufficient condition
for $CP$ violation in the standard model is then
${\cal J} \neq 0$ or
$\varphi \neq 0$, $\pi$.
Since $s_{\rm u} =0$ or $c_{\rm u} =0$ will definitely
(though indirectly) lead to $\varphi =0$
or $\pi$, it is unnecessary to count the condition
$\theta_{\rm u} \neq 0$ or $\pi/2$ together with
$\varphi \neq 0$ or $\pi$.
So is the situation for $\theta_{\rm d}$ and $\theta$.
In reality quark masses of each sector
have been found to perform a clear hierarchy, and all
elements of the flavor mixing matrix are nonvanishing \cite{PDG98}.
Therefore the realistic
condition for $CP$ violation is only associated with
the existence of one nontrivial phase parameter in $V$,
which in turn requires (at least) one nontrivial phase
difference
between $M_{\rm u}$ and $M_{\rm d}$.
\subsection{Exact analytical result for ${\cal J}$}
Let us derive the exact analytical relation
between the $CP$-violating parameter ${\cal J}$ and
the quark mass-eigenvalue differences.
Without loss of generality, we
just adopt the basis of flavor space which
accommodates the hermitian
quark mass matrices $M_{\rm u}$ and $M_{\rm d}$ in (4.1).
A basis-independent proof can similarly be carried out for two
arbitrary mass matrices $M'_{\rm u}$ and $M'_{\rm d}$, if one starts from
the hermitian products $H_{\rm u} \equiv M'_{\rm u} {M'_{\rm u}}^{\dagger}$
and $H_{\rm d} \equiv M'_{\rm d} {M'_{\rm d}}^{\dagger}$ and arranges them
to be of the same form as $M_{\rm u}$ and $M_{\rm d}$.
This can always be done by appropriately adjusting
the fields of right-handed quarks, which are iso-singlets in the
standard model.
For convenience we decompose $M_{\rm q}$ into
$M_{\rm q} = P_{\rm q}^{\dagger}
\overline{M}_{\rm q} P_{\rm q}$, where
\begin{equation}
\overline{M}_{\rm q} \; =\; \left ( \matrix{
E_{\rm q} & |D_{\rm q}| & {\bf 0} \cr
|D_{\rm q}| & C_{\rm q} & |B_{\rm q}| \cr
{\bf 0} & |B_{\rm q}| & A_{\rm q} \cr} \right ) \;
\end{equation}
is a real symmetric matrix, and
$P_{\rm q} = {\rm Diag} \{ 1, e^{{\rm i}\phi_{D_{\rm q}}},
e^{{\rm i} (\phi_{B_{\rm q}} + \phi_{D_{\rm q}})} \}$
is a diagonal phase matrix. In the following we shall
neglect the subscript ``q'', only if there is no
necessity to distinguish between the up and down quark sectors.
$\overline{M}$ can be diagonalized by use of the orthogonal transformation
$O^{\rm T} \overline{M} O = {\rm Diag} \{ \lambda_1, \lambda_2, \lambda_3 \}$,
where $\lambda_i$ (for $i=1,2,3$) are quark mass eigenvalues
and may be either positive or negative.
As a result, we have
\begin{eqnarray}
\sum_{i=1}^3 \lambda_i & = & A + C + E \; ,
\nonumber \\
\prod_{i=1}^3 \lambda_i & = & ACE - A |D|^2 - E |B|^2 \; ,
\nonumber \\
\sum_{i=1}^3 \lambda^2_i & = & A^2 + 2|B|^2 + C^2 + 2|D|^2 + E^2 \; .
\end{eqnarray}
It is a simple exercise to solve the nine matrix elements of
$O$ in terms of the parameters of quark mass matrices. Explicitly,
three diagonal elements of $O$ read
\footnote{Here and hereafter, the off-diagonal elements $B$ and $D$ are
both taken to be nonvanishing. The relevant calculations will somehow
be simplified if one of them vanishes.} :
\begin{eqnarray}
O_{11} & = & \left [ 1 + \left ( \frac{\lambda_1 - E}{|D|} \right )^2 + \left
( \frac{|B|}{|D|} \cdot \frac{\lambda_1 - E}{\lambda_1 - A} \right )^2
\right ]^{-1/2} \; , \nonumber \\
O_{22} & = & \left [ 1 + \left ( \frac{|D|}{\lambda_2 - E} \right )^2 + \left
( \frac{|B|}{\lambda_2 - A} \right )^2 \right ]^{-1/2} \; , \nonumber \\
O_{33} & = & \left [ 1 + \left ( \frac{\lambda_3 - A}{|B|} \right )^2 + \left
( \frac{|D|}{|B|} \cdot \frac{\lambda_3 - A}{\lambda_3 - E} \right )^2
\right ]^{-1/2} \; ;
\end{eqnarray}
and then six off-diagonal elements of $O$ can be obtained from the
relations
\begin{eqnarray}
O_{2i} & = & \frac{\lambda_i - E}{|D|} O_{1i} \; , \nonumber \\
O_{3i} & = & \frac{|B|}{\lambda_i - A}O_{2i} \; .
\end{eqnarray}
The flavor mixing matrix turns out to be $V \equiv O^{\rm T} (P_{\rm u}
P_{\rm d}^{\dagger}) O_{\rm d}$. More specifically, we have
\begin{equation}
V_{i\alpha} \; =\; O^{\rm u}_{1i} O^{\rm d}_{1\alpha} ~ + ~
O^{\rm u}_{2i} O^{\rm d}_{2\alpha} e^{{\rm i}\phi_1} ~ + ~
O^{\rm u}_{3i} O^{\rm d}_{3\alpha} e^{{\rm i}(\phi_1 + \phi_2)} \; ,
\end{equation}
where the Latin subscript $i$ and the Greek subscript
$\alpha$ run over $(u,c,t)$ and $(d,s,b)$ respectively,
and the phase differences $\phi_{1,2}$ have been
defined in (4.2).
The $CP$-violating parameter $\cal J$ can be calculated
from the common imaginary part of nine rephasing
invariants of $V$, i.e.,
${\cal J} = | {\rm Im} (V_{i\alpha} V_{j\beta} V^*_{i\beta}
V^*_{j\alpha}) |$ for $i\neq j$ and
$\alpha \neq \beta$ \cite{J85}.
With the help of (4.10) and (4.11), one
may express $\cal J$ in terms of
the parameters of quark mass matrices. After a lengthy calculation, we
arrive at
the following exact and rephasing-invariant result:
\begin{equation}
{\cal J} \; = \; \left (\lambda_i - \lambda_j \right )
\left ( \lambda_\alpha
- \lambda_\beta \right ) ~ f^{ij}_{\alpha\beta} \; ,
\end{equation}
where
\begin{eqnarray}
f^{ij}_{\alpha\beta} & = &
\frac{\left ( O^{\rm u}_{1i} O^{\rm u}_{1j}
O^{\rm d}_{1\alpha} O^{\rm d}_{1\beta} \right )^2}{|D_{\rm u}
D_{\rm d}|} ~
\left [ T_1 \sin \phi_1 ~ + ~ T_2 \sin\phi_2 ~ + ~
T_3 \sin (\phi_1 + \phi_2) \right . \nonumber \\
& & \left . + ~ T_4 \sin (\phi_1 - \phi_2) ~ + ~
T_5 \sin (2\phi_1 + \phi_2) ~ + ~ T_6 \sin (\phi_1 + 2\phi_2)
\right ] \; .
\end{eqnarray}
The expressions of $T_i$ (for $i=1,2,\cdot\cdot\cdot ,6$)
are listed in Appendix A.
One can see that $\cal J$ depends definitely on the mass-eigenvalue
differences $(\lambda_i - \lambda_j)$ of the up sector and
$(\lambda_\alpha - \lambda_\beta)$ of the down sector. Since the subscripts
$(i,j)$ and $(\alpha, \beta)$ run over the
corresponding quarks $(u,c,t)$ and $(d,s,b)$, $\cal J$ would vanish if
any two quarks with the same charge were degenerate in mass eigenvalues.
Therefore $\cal J$
carries the same information about $CP$ violation as ${\rm Det} ~\cal C$
in (4.6). Two remarks are in order.
(a) Note that a phase combination in the form of
$\sin (2\phi_1)$, $\sin (2\phi_2)$ or $\sin 2(\phi_1
+ \phi_2)$ has no contribution to $\cal J$. The reason
is simply that in $\cal J$ the terms associated with $e^{+{\rm i} 2\phi_1}$
and $e^{-{\rm i} 2\phi_1}$ have the same magnitude and cancel
each other. So it the case for the terms
associated with $e^{\pm {\rm i} 2\phi_2}$ and
$e^{\pm {\rm i} 2 (\phi_1 + \phi_2)}$.
Once the hierarchy of
$M_{\rm u}$ and $M_{\rm d}$ is taken into account, the magnitude of
$\cal J$ is expected to be dominated by the term proportional to $\sin\phi_1$
(see the next subsection).
(b) The dependence of $\cal J$ on the product of two mass-eigenvalue
differences
$(\lambda_i - \lambda_j)$ and $(\lambda_\alpha - \lambda_\beta)$
is indeed a basis-independent result, although we have obtained it in
a specific basis of flavor space for the quark mass matrices.
The basis-independent calculation of $\cal J$ is straightforward,
as we have mentioned above, if one starts from $H_{\rm u}
= M'_{\rm u} {M'_{\rm u}}^{\dagger}$ and
$H_{\rm d} = M'_{\rm d} {M'_{\rm d}}^{\dagger}$ for
arbitrary $M'_{\rm u}$ and $M'_{\rm d}$.
In this case it is easy to find
\begin{equation}
{\cal J} \; = \; \left (\lambda^2_i - \lambda^2_j \right )
\left ( \lambda^2_\alpha - \lambda^2_\beta \right ) ~
F^{ij}_{\alpha\beta}\; ,
\end{equation}
where $F^{ij}_{\alpha\beta}$ can be read off from $f^{ij}_{\alpha\beta}$
through the replacements of matrix elements from $M_{\rm u,d}$
to $H_{\rm u,d}$. Of course the results in (4.12) and
(4.14) essentially have the same physical meaning.
In reality it is known that quark masses show a strong
hierarchy in either sector. Therefore $\cal J$ vanishes
if and only if both $\phi_1 =0$ (or $\pi$) and $\phi_2 =0$ (or $\pi$)
hold.
The necessary and sufficient condition for $CP$ violation in the
standard model is trivially $\varphi \neq 0$ or $\pi$.
\subsection{Flavor mixing angles and the $CP$-violating phase}
Now let us take the hierarchy of quark masses
($|\lambda_1| \ll |\lambda_2| \ll |\lambda_3|$)
into account for the
hermitian mass matrices $M_{\rm u}$ and $M_{\rm d}$ in (4.1). This implies
$|A_{\rm q}| \gg |B_{\rm q}|, |C_{\rm q}| \gg |D_{\rm q}|,
|E_{\rm q}|$ for both sectors. Our purpose is to
calculate the
mixing angles ($\theta_{\rm u}$, $\theta_{\rm d}$ and $\theta$)
and the $CP$-violating phase ($\varphi$) in an analytically
exact way.
Certainly the orthogonal matrix $O$ used to diagonalize
$\overline{M}$ in (4.7) can further be written as a
product of three matrices $R_{12}$,
$R_{23}$ and $R_{31}$, which describe simple rotations in the 1--2,
2--3 and 3--1 planes respectively:
\begin{eqnarray}
R_{12}(\omega) & = & \left ( \matrix{
c_{\omega} & s_{\omega} & 0 \cr
- s_{\omega} & c_{\omega} & 0 \cr
0 & 0 & 1 \cr} \right ) \; , \nonumber \\ \nonumber \\
R_{23}(\sigma) & = & \left ( \matrix{
1 & 0 & 0 \cr
0 & c_{\sigma} & s_{\sigma} \cr
0 & - s_{\sigma} & c_{\sigma} \cr} \right ) \; , \nonumber \\
\nonumber \\
R_{31}(\tau) & = & \left ( \matrix{
c_{\tau} & 0 & s_{\tau} \cr
0 & 1 & 0 \cr
- s_{\tau} & 0 & c_{\tau} \cr} \right ) \; ,
\end{eqnarray}
where $s_{\omega} \equiv \sin \omega$, $c_{\omega} \equiv \cos
\omega$, etc. Taking $O= R_{13}R_{23}R_{12}$ for example, we arrive at
\begin{eqnarray}
\tan\tau & = & \frac{|D|}{|B|} \cdot \frac{(E-\lambda_1) +
(C-\lambda_2)}{\lambda_3 - E} \; , \nonumber \\
\tan\sigma & = & \frac{|D|}{|B|} \cdot \frac{(E-\lambda_1) +
(C-\lambda_2)}{\sqrt{|D|^2 + |B|^2 \tan^2\tau}} \; , \nonumber \\
\tan\omega & = & \frac{|D|}{\lambda_2 - E} \cdot
\frac{c^{~}_{\sigma}}{c^{~}_{\tau}} ~ + ~ s^{~}_{\sigma}
\tan\tau \; .
\end{eqnarray}
In view of the hierarchy of quark mass matrices, we find that the
magnitude of $\tan\tau$ is highly suppressed, leading to an excellent
approximation $\tau \approx 0^{\circ}$. Thus the matrix $O$ is
dominantly described by only two rotation angles, $\omega$ and
$\sigma$. This is naturally expected,
since in lowest order the diagonalization of the
mass matrices is provided by a rotation in the 2--3 plane and a rotation in
the 1--2 plane. Due to the vanishing of the (1,3) and (3,1) matrix
elements, a rotation in the 3--1 plane is
essentially unnecessary. This approximation has
been used to derive an interesting parametrization of the flavor
mixing matrix \cite{F79,Froggatt79,Hall}, whose form is quite similar to
that given in (2.2). Note, however, that the
exact parametrization (2.2) is indeed independent of the above
approximation, because the contribution from rotation matrices $R^{\rm
u}_{13}$ (up) and $R^{\rm d}_{13}$ (down)
to the flavor mixing matrix can always
be absorbed by redefining its three overall mixing angles. Since the
concrete calculation of those mixing angles from $R^{\rm u,d}_{12}$,
$R^{\rm u,d}_{23}$ and $R^{\rm u,d}_{31}$ is rather complicated
and less instructive (see,
e.g., Ref. \cite{Rasin}), we shall subsequently follow a different and
more straightforward procedure towards the same goal.
We make use of the expression of $V$ given in (4.11).
The parametrization of $V$ in
terms of three mixing angles ($\theta_{\rm u}$, $\theta_{\rm d}$,
$\theta$) and one $CP$-violating phase ($\varphi$) has been shown in
(2.2). To link these four parameters with the parameters of
quark mass matrices in a concise way, we first define four dimensionless
quantities:
\begin{eqnarray}
X_{\rm u} & \equiv & \left | \frac{|D_{\rm u}|}{\lambda^{\rm u}_1 -
E_{\rm u}} \cdot
\frac{|D_{\rm d}| \left (\lambda^{\rm d}_3 - A_{\rm d} \right )}
{|B_{\rm d}| \left
(\lambda^{\rm d}_3 - E_{\rm d} \right )} ~ + ~ \frac{\lambda^{\rm d}_3
- A_{\rm d}}{|B_{\rm d}|}
e^{{\rm i}\phi_1} ~ + ~ \frac{|B_{\rm u}|}{\lambda^{\rm u}_1 - A_{\rm u}}
e^{{\rm i}(\phi_1 + \phi_2)}
\right | \; , \nonumber \\
Y_{\rm u} & \equiv & \left | \frac{|D_{\rm u}|}{\lambda^{\rm u}_2 -
E_{\rm u}} \cdot
\frac{|D_{\rm d}| \left (\lambda^{\rm d}_3 - A_{\rm d} \right )}
{|B_{\rm d}| \left
(\lambda^{\rm d}_3 - E_{\rm d} \right )} ~ + ~
\frac{\lambda^{\rm d}_3 - A_{\rm d}}{|B_{\rm d}|}
e^{{\rm i}\phi_1} ~ + ~ \frac{|B_{\rm u}|}{\lambda^{\rm u}_2 - A_{\rm u}}
e^{{\rm i}(\phi_1 + \phi_2)}
\right | \; ;
\end{eqnarray}
and $(X_{\rm d}, Y_{\rm d})$ can directly be obtained from $(X_{\rm
u}, Y_{\rm u})$ through the subscript exchange ${\rm u}
\leftrightarrow {\rm d}$ in (4.17).
After a lengthy but straightforward calculation, we arrive at
\begin{eqnarray}
\tan\theta_{\rm u} & = & \frac{O^{\rm u}_{21}}{O^{\rm u}_{22}} \cdot
\frac{X_{\rm u}}{Y_{\rm u}} \; , \nonumber \\
\tan\theta_{\rm d} & = & \frac{O^{\rm d}_{21}}{O^{\rm d}_{22}} \cdot
\frac{X_{\rm d}}{Y_{\rm d}} \; ,
\end{eqnarray}
and
\begin{eqnarray}
\sin\theta & = & \left [ (O^{\rm u}_{21} )^2 X^2_{\rm u}
~ + ~ (O^{\rm u}_{22} )^2 Y^2_{\rm u} \right ]^{1/2}
O^{\rm d}_{33} \; , \nonumber \\
& = & \left [ (O^{\rm d}_{21} )^2 X^2_{\rm d}
~ + ~ (O^{\rm d}_{22} )^2 Y^2_{\rm d} \right ]^{1/2}
O^{\rm u}_{33} \; ,
\end{eqnarray}
where $O_{21}$, $O_{22}$ and $O_{33}$ for up and down sectors have been
given in (4.9) and (4.10). Also an indirect relation
between $\varphi$ and $\phi_{1,2}$ can be obtained as follows:
\begin{equation}
\cos \varphi \; =\; \frac{s^2_{\rm u} c^2_{\rm d} c^2 + c^2_{\rm u}
s^2_{\rm d} - |V_{us}|^2}{2 s_{\rm u} c_{\rm u} s_{\rm d} c_{\rm d} c}
\; ,
\end{equation}
where
\begin{equation}
|V_{us}| \; =\; O^{\rm u}_{11} O^{\rm d}_{22} \left | \frac{|D_{\rm
d}|}{\lambda^{\rm d}_2 - E_{\rm d}} + \frac{\lambda^{\rm u}_1 -
E_{\rm u}}{|D_{\rm u}|}
e^{{\rm i} \phi_1} \left ( 1 +
\frac{|B_{\rm u}|}{\lambda^{\rm u}_1 - A_{\rm u}} \cdot \frac{|B_{\rm
d}|}{\lambda^{\rm d}_2 - A_{\rm d}} e^{{\rm i} \phi_2} \right ) \right | \; .
\end{equation}
If the hierarchies of the matrix elements and mass eigenvalues of
$M_{\rm u, d}$ are taken into account, one can see that the
effect of $\phi_2$ on $|V_{us}|$ is strongly suppressed and
thus negligible.
Fitting $|V_{us}|$ with current data should essentially determine the
magnitude of $\phi_1$. Note also that the terms associated with $\phi_1$ and
$\phi_2$ may primarily be cancelled in the ratios $X_{\rm u}/Y_{\rm u}$
and $X_{\rm d}/Y_{\rm d}$ due to the hierarchical structures of $M_{\rm u}$
and $M_{\rm d}$, hence the dependence of $\theta_{\rm u}$ and $\theta_{\rm d}$
on $\phi_{1,2}$ could be negligible in the leading order approximation.
Although the mixing angle $\theta$ may be sensitive to $\phi_1$ and $\phi_2$
(or one of them), its smallness indicated by current data makes the
factor $\cos\theta$
in the denominator of $\cos\varphi$ completely negligible. As a result,
(4.20) and (4.21)
imply that the $CP$-violating phase $\varphi$ depends dominantly on $\phi_1$
through $|V_{us}|$, unless the magnitude of $\phi_1$ is very small.
Without fine-tuning, we find that a delicate numerical
analysis does support the argument made here, i.e.,
$\phi_2$ plays a negligible role for $CP$ violation in $V$,
because of the hierarchy of quark masses.
The observed $CP$ violation is
linked primarily to the phases in the (1,2) and
(2,1) elements of the quark mass matrices.
\section{A realistic texture of mass matrices}
\setcounter{equation}{0}
In order to get definite predictions for the flavor mixing angles and $CP$
violation, we proceed to specify the general hermitian mass matrices
in (2.1) or (4.1) by taking $E_{\rm q} =0$:
\begin{equation}
M_{\rm q} \; = \; \left ( \matrix{
{\bf 0} & D_{\rm q} & {\bf 0} \cr
D^*_{\rm q} & C_{\rm q} & B_{\rm q} \cr
{\bf 0} & B^*_{\rm q} & A_{\rm q} \cr } \right ) \; .
\end{equation}
In case of two quark families, this is just the
form taken for $\tilde{M}_{\rm d}$ in (3.7).
As remarked above, the texture zeros
in (1,3) and (3,1) positions can always be arranged.
Thus the physical constraint is as follows: in the
flavor basis in which (1,3) and (3,1) elements of $M_{\rm u,d}$
vanish, the (1,1) element of $M_{\rm u,d}$ vanishes as well.
This can strictly be true only at a particular energy scale. The vanishing of
the (1,1) element can be viewed as a result of
an underlying flavor symmetry, which
may either be discrete or continuous. In the literature a number of such
possibilities have been discussed (see, e.g.,
Refs. \cite{F79} -- \cite{Zero4}).
Here we shall not discuss further details in this respect,
but concentrate on the phenomenological
consequences of such a texture pattern.
It is particularly interesting that some predictions of this ansatz
for the mixing angles and the unitarity triangle are approximately
independent of the renormalization-group effects, therefore
a specification of the energy scale at which the texture of
$M_{\rm u,d}$ holds is unnecessary for our purpose.
We believe that $M_{\rm q}$ given in (5.1) is
a realistic candidate for the quark mass matrices of a
(yet unknown) fundamental theory responsible for fermion mass generation
and $CP$ violation, and we shall make some further speculations
about this point at the end of this paper.
\subsection{Flavor mixing angles}
We take $C_{\rm q} \neq 0$ and define
$|B_{\rm q}|/C_{\rm q} \equiv r^{~}_{\rm q}$ for each
quark sector
\footnote{Note that the special condition $C_{\rm q}/A_{\rm q}
= |D_{\rm q}/B_{\rm q}|^2$ has been imposed on $M_{\rm q}$ in
a recent paper \cite{Froggatt}. It leads to
vanishing flavor mixing among all three quark families
in the chiral limit of $u$- and $d$-quark
masses, and in turn requires a kind of correlation between the flavor
mixing angles $\theta_{\rm u}$, $\theta_{\rm d}$ and $\theta$.
This unusual feature is apparently in conflict with our
arguments made in section 3 (see also Ref. \cite{F87}).
Beyond the chiral symmetry limit the aforementioned condition
is equivalent to taking
$|r_{\rm u}| \approx \sqrt{m_um_t}/m_c$ and
$|r_{\rm d}| \approx \sqrt{m_dm_b}/m_s$
in our case. Clearly both
$r_{\rm u}$ and $r_{\rm d}$ are of $O(1)$ in magnitude, consistent with
the common expectation.}.
The magnitude of $r^{~}_{\rm q}$
is expected to be of $O(1)$.
The parameters $A_{\rm q}$, $|B_{\rm q}|$, $C_{\rm q}$
and $|D_{\rm q}|$ in (5.1)
can be expressed in terms of the quark mass eigenvalues and $r^{~}_{\rm q}$.
Applying such results to the general formulas listed in (4.17) --
(4.19), we get three mixing angles of $V$ as follows:
\begin{eqnarray}
\tan\theta_{\rm u} & = & \sqrt{\frac{m_u}{m_c}} ~ \left ( 1 + \Delta_{\rm u}
\right ) \; , \nonumber \\
\tan\theta_{\rm d} & = & \sqrt{\frac{m_d}{m_s}} ~ \left ( 1 + \Delta_{\rm d}
\right ) \; , \nonumber \\
\sin\theta & = & \left | r_{\rm d} \frac{m_s}{m_b} \left (1 -
\delta_{\rm d} \right ) ~ - ~ r_{\rm u} \frac{m_c}{m_t} \left ( 1 -
\delta_{\rm u} \right ) e^{{\rm i}\phi_2} \right | \; ,
\end{eqnarray}
where the next-to-leading order corrections read
\begin{eqnarray}
\Delta_{\rm u} & = & \sqrt{\frac{m_c m_d}{m_u m_s}} ~ \frac{m_s}{m_b}
~ \left | {\rm Re} \left [ e^{{\rm i}\phi_1} ~ - ~ \frac{r_{\rm
u}}{r_{\rm d}} \cdot \frac{m_c m_b}{m_t m_s}
e^{{\rm i}(\phi_1 + \phi_2)} \right
]^{-1} \right | \; , \nonumber \\
\Delta_{\rm d} & = & \sqrt{\frac{m_u m_s}{m_c m_d}} ~ \frac{m_c}{m_t}
~ \left | {\rm Re} \left [ e^{{\rm i}\phi_1} ~ - ~ \frac{r_{\rm
d}}{r_{\rm u}} \cdot \frac{m_t m_s}{m_c m_b}
e^{{\rm i}(\phi_1 + \phi_2)} \right
]^{-1} \right | \; ;
\end{eqnarray}
and
\begin{eqnarray}
\delta_{\rm u} & = & \frac{m_u}{m_c} ~ +
\left ( 1 + r^2_{\rm u} \right ) \frac{m_c}{m_t}
\; , \nonumber \\
\delta_{\rm d} & = & \frac{m_d}{m_s} ~ + \left ( 1 + r^2_{\rm d}
\right ) \frac{m_s}{m_b}
\; .
\end{eqnarray}
Clearly the result for $\hat{\delta}_{\rm u,d}$ in (3.4) can be
reproduced from $\delta_{\rm u,d}$ in (5.4), if one takes the chiral
limit $m_u \rightarrow 0$, $m_d \rightarrow 0$.
From (5.2) we also observe that the phase $\phi_2$ is
only associated with the small quantity $m_c/m_t$ in $\sin\theta$.
To get the relationship between $\varphi$ and $\phi_1$ or $\phi_2$,
we first calculate $|V_{us}|$ from the quark mass matrices by use of
(4.21). It turns out that
\begin{equation}
|V_{us}| \; = \; \left (1 -\frac{1}{2} \frac{m_u}{m_c} - \frac{1}{2}
\frac{m_d}{m_s} \right ) \left | \sqrt{\frac{m_d}{m_s}} ~ - ~
\sqrt{\frac{m_u}{m_c}} ~ e^{{\rm i}\phi_1} \right | \;
\end{equation}
in the next-to-leading order approximation. Note that this result can
also be achieved from (3.8) and (3.9), which were obtained in the
heavy quark limit. Confronting (5.5) with current data on
$|V_{us}|$ leads to the result $\phi_1 \sim
90^{\circ}$, as we have discussed before. Therefore $\cos\phi_1$ is
expected to be a small quantity.
Then we use (4.20) together with (5.2) and (5.5)
to calculate $\cos\varphi$.
In the same order approximation, we arrive at
\begin{equation}
\cos\varphi \; =\; \sqrt{\frac{m_u m_s}{m_c m_d}} ~ \Delta_{\rm u} ~ + ~
\sqrt{\frac{m_c m_d}{m_u m_s}} ~ \Delta_{\rm d} ~ + ~ \left (1 -
\Delta_{\rm u} - \Delta_{\rm d} \right ) \cos\phi_1 \; .
\end{equation}
The contribution of $\phi_2$ to $\varphi$ is substantially suppressed at
this level of accuracy.
For simplicity, we proceed by taking $r_{\rm u} = r_{\rm d} \equiv r$,
which holds in some models with natural flavor
symmetries \cite{Democracy}. Then
$\sin\theta$ becomes proportional to a universal parameter $|r|$.
In view of the fact $m_s/m_b \sim O(10) ~ m_c/m_t$, we find that the result
in (5.3) can be simplified as
\begin{eqnarray}
\Delta_{\rm u} & = & \sqrt{\frac{m_c m_d}{m_u m_s}} ~ \frac{m_s}{m_b}
~ \cos\phi_1 \; , \nonumber \\
\Delta_{\rm d} & = & 0 \; .
\end{eqnarray}
Also the relation between $\varphi$ and $\phi_1$ in (5.6) is simplified to
\begin{equation}
\cos\varphi \; =\; \left (1 + \frac{m_s}{m_b} \right ) \cos\phi_1 \; .
\end{equation}
As $m_s/m_b \sim 4\%$, it becomes apparent that
$\varphi \approx \phi_1$ is a good approximation.
Note that $\phi_1 = \varphi$ holds exactly in the
heavy quark limit, in which $\varphi$ has been denoted
as $\tilde{\varphi}$ (see (3.5) as well as Fig. 1).
The equality $\phi_1 = \tilde{\varphi}$ follows,
i.e., both stand for the
phase difference between the mass matrix elements
$D_{\rm u}$ and $D_{\rm d}$.
Following (4.12) we evaluate the dependence of the
$CP$-violating measurable $\cal J$ on $\phi_1$, $\phi_2$
and their various combinations. The results for six
coefficients of $f^{ij}_{\alpha\beta}$ are listed in
Appendix B. We confirm that the magnitude of $\cal J$
is dominated by the $\sin\phi_1$ term and receives
one-order smaller corrections from the $\sin (\phi_1 \pm \phi_2)$
terms. As a result,
\begin{equation}
{\cal J} \; \approx \; |r|^2 \sqrt{\frac{m_u}{m_c}}
\sqrt{\frac{m_d}{m_s}} \left (\frac{m_s}{m_b}\right )^2
\sin\phi_1 \;
\end{equation}
holds to a good degree of accuracy. Clearly
${\cal J} \sim O(10^{-5}) \times \sin\phi_1$ with $\sin\phi_1 \sim 1$
is favored by current data.
The result of $\cal J$ in (5.9) might give the impression that
$CP$ violation is absent if either $m_u$ or $m_d$ vanishes. This
is not exactly true, however. If we set $m_u=0$,
$\cal J$ is not zero, but it becomes dependent on $\sin \phi_2$
with a factor which is about two orders of magnitude
smaller (i.e., of order $10^{-7}$):
\begin{equation}
{\cal J} \; \approx \; |r|^2 ~ \frac{m_c}{m_t} \cdot
\frac{m_d}{m_s} \left (\frac{m_s}{m_b} \right )^2
\sin\phi_2 \; .
\end{equation}
Certainly this possibility is already ruled out by experimental data.
Note also that the model predicts
\begin{eqnarray}
\tan\theta_{\rm u} \; =\;
\left | \frac{V_{ub}}{V_{cb}} \right | & = &
\sqrt{\frac{m_u}{m_c}} ~ \left (1 + \Delta_{\rm u} \right ) \; ,
\nonumber \\
\tan\theta_{\rm d} \; =\;
\left | \frac{V_{td}}{V_{ts}} \right | & = &
\sqrt{\frac{m_d}{m_s}} ~ \left (1 + \Delta_{\rm d} \right ) \; ,
\end{eqnarray}
a result obtained first by one of us from a
more specific pattern of quark mass matrices \cite{F78}.
In $B$-meson physics, $|V_{ub}/V_{cb}|$ can be determined
from the ratio of the decay rate of $B\rightarrow
(\pi, \rho) l \nu^{~}_l$ to that of $B\rightarrow D^* l\nu^{~}_l$;
and $|V_{td}/V_{ts}|$ can be extracted from the ratio of
the rate of $B^0_d$-$\bar{B}^0_d$ mixing to that of
$B^0_s$-$\bar{B}^0_s$ mixing.
\subsection{The unitarity triangle}
\begin{figure}[t]
\begin{picture}(400,160)(130,210)
\put(300,300){\line(1,0){150}}
\put(300,300.5){\line(1,0){150}}
\put(370,285.5){\makebox(0,0){$|V_{cd}|$}}
\put(300,300){\line(1,3){21.5}}
\put(300,300.5){\line(1,3){21.5}}
\put(300,299.5){\line(1,3){21.5}}
\put(292,333){\makebox(0,0){$S_{\rm u}$}}
\put(450,300){\line(-2,1){128}}
\put(450,300.5){\line(-2,1){128}}
\put(395,343.5){\makebox(0,0){$S_{\rm d}$}}
\put(315,310){\makebox(0,0){$\gamma$}}
\put(408,309){\makebox(0,0){$\beta$}}
\put(328,350){\makebox(0,0){$\alpha$}}
\end{picture}
\vspace{-2.6cm}
\caption{The rescaled unitarity triangle (UT) in the complex plane.}
\end{figure}
We are now in a position to calculate the unitarity triangle (UT) of
quark flavor mixing defined in (2.8), whose three inner angles are
denoted as $\alpha$, $\beta$ and $\gamma$ in (2.9). Note that three
sides of the unitarity triangle can be rescaled by $V_{cb}^*$ (see
Fig. 2 for illustration). The resultant triangle reads
\begin{equation}
|V_{cd}| \; =\; \left | S_{\rm d} ~ -~ S_{\rm u} ~ e^{-{\rm i}\alpha}
\right | \; ,
\end{equation}
where $S_{\rm u} = |V_{ub}V_{ud}/V_{cb}|$ and $S_{\rm d} =
|V_{tb}V_{td}/V_{cb}|$. After some calculations $S_{\rm u}$, $S_{\rm
d}$ and $\alpha$ are obtained from the above quark mass texture in the
next-to-leading order approximation:
\begin{eqnarray}
S_{\rm u} & = & \sqrt{\frac{m_u}{m_c}} \left ( 1 - \frac{1}{2}
\frac{m_u}{m_c} - \frac{1}{2} \frac{m_d}{m_s} + \sqrt{\frac{m_c
m_d}{m_u m_s}} ~ \frac{m_s}{m_b} ~ \cos\phi_1 + \sqrt{\frac{m_u
m_d}{m_c m_s}} ~ \cos\phi_1 \right ) \; , \nonumber \\ S_{\rm d} & = &
\sqrt{\frac{m_d}{m_s}} \left ( 1 + \frac{1}{2} \frac{m_u}{m_c} -
\frac{1}{2} \frac{m_d}{m_s} \right ) \; ;
\end{eqnarray}
and
\begin{equation}
\sin\alpha \; = \; \left ( 1 - \sqrt{\frac{m_u m_d}{m_c m_s}} ~
\cos\phi_1 \right ) \sin\phi_1 \; .
\end{equation}
A comparison of the rescaled UT in Fig. 2 with the LT in Fig. 1, which
is obtained in the heavy quark limit, is interesting. We find
\begin{eqnarray}
\frac{S_{\rm u} - R_{\rm u}}{R_{\rm u}} & = & \left ( 1 + \frac{m_c
m_s}{m_u m_b} \right ) \sqrt{\frac{m_u m_d}{m_c m_s}} ~
\cos\tilde{\varphi} \; , \nonumber \\ \frac{S_{\rm d} - R_{\rm
d}}{R_{\rm d}} & = & \frac{m_u}{m_c} \; , \nonumber \\
\frac{\sin\alpha - \sin\tilde{\varphi}}{\sin\tilde{\varphi}} & = & -
\sqrt{\frac{m_u m_d}{m_c m_s}} ~ \cos\tilde{\varphi} \; ,
\end{eqnarray}
which are of order $15\% \cos\tilde{\varphi}$, $0.4\%$ and $1.4\%
\cos\tilde{\varphi}$, respectively. Obviously $R_{\rm d} \approx
S_{\rm d}$ is an excellent approximation, and $\alpha \approx
\tilde{\varphi} \approx \varphi$ is a good approximation. As $\varphi$
(or $\tilde{\varphi}$) is expected to be close to $90^{\circ}$,
$R_{\rm u} \approx S_{\rm u}$ should also be accurate enough in the
next-to-leading order estimation. Therefore the light-quark triangle
is essentially {\it congruent with} the rescaled unitarity triangle!
This result has two straightforward implications: first, $CP$
violation is an effect arising primarily from the light quark sector;
second, the $CP$-violating observables ($\alpha$, $\beta$, $\gamma$)
can be predicted in terms of the light quark masses and the phase
difference between up and down mass matrices \cite{FX95}. If we use
the value of $|V_{cd}|$, which is expected to equal $|V_{us}|$ within
the $0.1\%$ error bar \cite{Xing96}, then all three angles of the
unitarity triangle can be calculated in terms of $m_u/m_c$, $m_d/m_s$
and $|V_{cd}|$ to a good degree of accuracy.
The three angles of the UT ($\alpha$, $\beta$ and $\gamma$) will be
well determined at the $B$-meson factories,
e.g., from the $CP$
asymmetries in $B_d\rightarrow \pi^+\pi^-$, $B_d\rightarrow J/\psi
K_S$ and $B^{\pm}_u\rightarrow (D^0, \bar{D}^0) + K^{(*)\pm}$
decays \cite{BB}.
The characteristic measurable quantities are $\sin
(2\alpha)$, $\sin (2\beta)$ and $\sin^2\gamma$, respectively. For
the purpose of illustration, we typically take
$|V_{us}| = |V_{cd}| =0.22$,
$m_u/m_c =0.0056$, $m_d/m_s = 0.045$ and $m_s/m_b = 0.033$ to
calculate these three $CP$-violating parameters from
the LT and from the rescaled UT separately.
Both approaches lead to
$\alpha \approx 90^{\circ}$,
$\beta \approx 20^{\circ}$ and
$\gamma \approx 70^{\circ}$,
which are in good agreement with the results obtained from
the standard
analysis of current data on $|V_{ub}/V_{cb}|$, $\epsilon^{~}_K$,
$B^0_d$-$\bar{B}^0_d$ mixing and $B^0_s$-$\bar{B}^0_s$ mixing
\cite{PRS98}. Note that among three $CP$-violating observables
only $\sin (2\beta)$ is remarkably sensitive to the value of
$m_u/m_c$, which involves quite large uncertainty (e.g., $\sin
(2\beta)$ may change from $0.4$ to $0.8$ if $m_u/m_c$ varies in the
range $0.002 - 0.01$). For this reason
we emphasize again that the numbers given above can only serve
as an illustration.
A more reliable determination of the
quark mass values is crucial, in order to test the ans$\rm\ddot{a}$tze
of quark mass matrices in a numerically decisive way
\footnote{A similar remark based on more delicate numerical
analysis has also be made in Ref. \cite{Barbieri}.}.
It is also worth mentioning that
the result $\tan\theta_{\rm d} = \sqrt{m_d/m_s}$ is particularly
interesting for the mixing rates of
$B^0_d$-$\bar{B}^0_d$ and $B^0_s$-$\bar{B}^0_s$
systems, measured by $x_{\rm d}$ and $x_{\rm s}$ respectively \cite{PDG98}.
The ratio $x_{\rm s} /x_{\rm d}$ amounts to $|V_{ts}/V_{td}|^2 =
\tan^{-2} \theta_{\rm d}$ multiplied by a factor $\chi_{\rm su(3)} = 1.45
\pm 0.13$, which reflects the $\rm SU(3)_{\rm flavor}$ symmetry breaking
effects \cite{Marti}. As $x_{\rm d} = 0.723 \pm 0.032$ has been well
determined \cite{PDG98}, the prediction for the value of $x_{\rm s}$ is
\begin{equation}
x_{\rm s} \; =\; x_{\rm d} ~ \chi_{\rm su(3)} ~ \frac{m_s}{m_d} \; =\;
19.8 \pm 3.5 \; ,
\end{equation}
where $m_s/m_d = 18.9 \pm 0.8$, obtained from the chiral perturbation
theory \cite{Leut}, has been used. This result is certainly consistent
with the present experimental bound on $x_{\rm s}$, i.e.,
$x_{\rm s} > 14.0$ at the $95\%$ confidence level \cite{PDG98}.
A measurement
of $x_{\rm s} \sim 20$ may be realized at the forthcoming
HERA-$B$ and LHC-$B$ experiments.
\subsection{Comparison with the Ramond-Roberts-Ross patterns}
The quark mass matrices $M_{\rm u}$ and $M_{\rm d}$ given in (5.1)
have parallel structures with four texture zeros (here a pair of
off-diagonal texture zeros are counted as one zero due to the
hermiticity of $M_{\rm u}$ and
$M_{\rm d}$). Giving up the parallelism between the
structures of $M_{\rm u}$ and $M_{\rm d}$, Ramond, Roberts and Ross
(RRR) have found that there exist five phenomenologically allowed
patterns of quark mass matrices -- each of them has five texture zeros
\cite{RRR},
as listed in Table 1. The RRR patterns I, II or IV can be formally regarded
as a special case of our four-texture-zero pattern (5.1), with
$B_{\rm u} =0$, $C_{\rm u} =0$ or $B_{\rm d} =0$, respectively. Note that
$M_{\rm u}$ of the RRR pattern III or V has nonvanishing (1,3) and (3,1)
elements, therefore these two patterns are essentially different
from the mass matrices assumed
in (4.1) or (5.1). As a comparison, here
we make some brief comments on consequences of the RRR patterns on the
flavor mixing angles $\theta$, $\theta_{\rm u}$ and $\theta_{\rm d}$.
\scriptsize
\begin{table}[t]
\caption{Five RRR patterns of quark mass matrices.}
\vspace{0.2cm}
\begin{center}
\begin{tabular}{ccccccc}\hline\hline
Pattern & I & II & III & IV & V \\ \hline \\
$M_{\rm u}$ & $\left ( \matrix{ 0 & D_{\rm u} & 0 \cr
D^*_{\rm u} & C_{\rm u} & 0 \cr
0 & 0 & A_{\rm u} \cr } \right )$
& $\left ( \matrix{ 0 & D_{\rm u} & 0 \cr
D^*_{\rm u} & 0 & B_{\rm u} \cr
0 & B^*_{\rm u} & A_{\rm u} \cr } \right )$
& $\left ( \matrix{ 0 & 0 & F_{\rm u} \cr
0 & C_{\rm u} & 0 \cr
F^*_{\rm u} & 0 & A_{\rm u} \cr } \right )$
& $\left ( \matrix{ 0 & D_{\rm u} & 0 \cr
D^*_{\rm u} & C_{\rm u} & B_{\rm u} \cr
0 & B^*_{\rm u} & A_{\rm u} \cr } \right )$
& $\left ( \matrix{ 0 & 0 & F_{\rm u} \cr
0 & C_{\rm u} & B_{\rm u} \cr
F^*_{\rm u} & B_{\rm u}^* & A_{\rm u} \cr } \right )$ \\ \\
$M_{\rm d}$ & $\left ( \matrix{ 0 & D_{\rm d} & 0 \cr
D^*_{\rm d} & C_{\rm d} & B_{\rm d} \cr
0 & B^*_{\rm d} & A_{\rm d} \cr } \right )$
& $\left ( \matrix{ 0 & D_{\rm d} & 0 \cr
D^*_{\rm d} & C_{\rm d} & B_{\rm d} \cr
0 & B^*_{\rm d} & A_{\rm d} \cr } \right )$
& $\left ( \matrix{ 0 & D_{\rm d} & 0 \cr
D^*_{\rm d} & C_{\rm d} & B_{\rm d} \cr
0 & B^*_{\rm d} & A_{\rm d} \cr } \right )$
& $\left ( \matrix{ 0 & D_{\rm d} & 0 \cr
D_{\rm d}^* & C_{\rm d} & 0 \cr
0 & 0 & A_{\rm d} \cr } \right )$
& $\left ( \matrix{ 0 & D_{\rm d} & 0 \cr
D^*_{\rm d} & C_{\rm d} & 0 \cr
0 & 0 & A_{\rm d} \cr } \right )$ \\ \\ \hline \hline
\end{tabular}
\end{center}
\end{table}
\normalsize
(a) For the RRR pattern I, the magnitude of $\sin\theta$ is governed by
the (2,3) and (3,2) elements of $M_{\rm d}$. Therfore $|B_{\rm d}| \sim
|C_{\rm d}|$ is expected, to result in $\sin\theta \sim m_s/m_b$. In
the leading order approximation, $\tan\theta_{\rm u} = \sqrt{m_u/m_c}$
and $\tan\theta_{\rm d} = \sqrt{m_d/m_s}$ hold. The next-to-leading
order corrections to these two quantities are almost indistinguishable
from those obtained in (5.3) and (5.7) for our four-texture-zero
ansatz.
(b) The mass matrix $M_{\rm u}$ of the RRR pattern II takes the
well-known form suggested originally by one of us in Refs.
\cite{F79,F77}. To reproduce the experimental
value of $\sin\theta$, the possibility $|B_{\rm d}| \gg |C_{\rm d}|$
has been abandoned and the condition $|B_{\rm d}| \sim |C_{\rm d}|$
is required. However, significant cancellation between the term
proportional to $\sqrt{m_c/m_t}$ (from $M_{\rm u}$) \cite{F79}
and that proportional to $m_s/m_b$ (from $M_{\rm d}$) in
$\sin\theta$ may take place, if the phase difference between
$B_{\rm u}$ and $B_{\rm d}$ is vanishing or very small. The leading
order results $\tan\theta_{\rm u} = \sqrt{m_u/m_c}$ and
$\tan\theta_{\rm d} = \sqrt{m_d/m_s}$ can still be obtained here,
but their next-to-leading order corrections may deviate somehow
from those obtained in the above subsections.
(c) From the RRR pattern IV one can arrive at $\sin\theta \sim \sqrt{m_c/m_t}$
with the necessary condition $|C_{\rm u}| \ll |B_{\rm u}|$, since the
(2,3) and (3,2) elements of $M_{\rm d}$ vanish. The results for
$\tan\theta_{\rm u}$ and $\tan\theta_{\rm d}$ are similar to those
obtained from the RRR pattern I.
(d) The nonvanishing (1,3) and (3,1) elements of $M_{\rm u}$ in the
RRR pattern III make its prediction for the mixing angles $\theta_{\rm u}$
and $\theta_{\rm d}$ quite different from all patterns discussed above.
Analytically one can find $\tan\theta_{\rm u} \sim (m_b/m_s) \sqrt{m_u/m_t}$ ,
while $\tan\theta_{\rm d}$ is a complicated combination of the
terms $\sqrt{m_d/m_s}$ and $(m_b/m_s) \sqrt{m_u/m_t}$ with a relative
phase.
In addition, $\sin\theta \sim m_s/m_b$ holds under the condition
$|B_{\rm d}| \sim |C_{\rm d}|$, similar to the RRR pattern I.
(e) For the RRR pattern V, the necessary condition $|B_{\rm u}| \gg
|C_{\rm u}|$ is required in order to reproduce
$\sin\theta \sim \sqrt{m_c/m_t}$ .
Here again the nonvanishing (1,3) and (3,1) elements of $M_{\rm u}$
result in very complicated expressions for $\tan\theta_{\rm u}$ and
$\tan\theta_{\rm d}$
(even more complicated than those in the RRR pattern III \cite{KX98}).
For reasons of naturalness and simplicity, we argue that
the RRR patterns III and V are unlikely to be good candidates for the
quark mass matrices in an underlying theory of fermion mass generation.
\section{Discussions and conclusion}
\setcounter{equation}{0}
We have studied the phenomena of quark flavor mixing and $CP$
violation in the context of generic hermitian mass matrices.
The necessary and sufficient conditions for $CP$ violation in
the standard model
have been clarified at both the level of quark mass matrices
and that of the flavor mixing matrix. Our particular observation
is that $CP$ violation is primarily linked to a phase
difference of about $90^{\circ}$ in the light quark sector,
and this property becomes most apparent in the new
parametrization (2.2). To be more
specific, we have analyzed a realistic pattern of quark mass
matrices with four texture zeros and given predictions
for the flavor mixing and $CP$-violating parameters.
The approximate congruency between the light-quark triangle (LT)
and the rescaled unitarity triangle (UT), which
provides an intuitive and scale-independent connection of
$CP$-violating observables to quark mass ratios, is
particularly worth mentioning.
Let us make some further comments on the quark mass matrix
(5.1), its phenomenological hints and its theoretical
prospects.
Naively one might not expect any prediction from the
four-texture-zero mass matrices in (5.1),
since they totally consist of ten free parameters
(two of them are the phase differences between $M_{\rm u}$ and $M_{\rm d}$).
This is not true, however, as we have seen.
We find that two predictions,
$\tan\theta_{\rm u} \approx \sqrt{m_u/m_c}$ and $\tan\theta_{\rm d} \approx
\sqrt{m_d/m_s}$ , can be obtained in the leading order approximation.
In some cases
the latter may even hold in the next-to-leading order
approximation, as shown in (5.2) and (5.7).
Note again that these two relations,
as a consequence of the hierarchy and texture zeros of our
quark mass matrices, are essentially independent of the renormalization-group
effects. This interesting scale-independent
feature can also be seen from the LT and the rescaled
UT as well as their
inner angles $(\alpha, \beta, \gamma)$.
It remains to be seen whether the interesting possibility
$\varphi \approx \phi_1 \approx 90^{\circ}$, indicated by current
data of quark masses and flavor mixing, could arise from an
underlying flavor symmetry or a dynamical
symmetry breaking scheme. Some speculations about this problem
have been made
(see, e.g., Refs. \cite{FX95,Weyers} and Refs. \cite{Democracy,New}).
However, no final conclusion has been reached thus far.
It is remarkable, nevertheless, that we have at least observed a
useful relation between the area of the UT
(${\cal A}_{\rm UT}$) and that of the LT
(${\cal A}_{\rm LT}$) to a good degree of accuracy:
\begin{equation}
{\cal A}_{\rm UT} \; \approx \; |V_{cb}|^2 {\cal A}_{\rm LT}
\; \approx \; \sin^2\theta ~ {\cal A}_{\rm LT} \; .
\end{equation}
Since ${\cal A}_{\rm UT} = {\cal J}/2$ measures the magnitude
of $CP$ violation in the standard model, we conclude that
$CP$ violation is primarily linked to the
light quark sector. This is a natural consequence of the
strong hierarchy between the heavy and light quark masses,
which is on the other hand responsible for the smallness of
${\cal J}$ or ${\cal A}_{\rm UT}$.
Is it possible to derive the quark mass matrix (5.1) in some
theoretical frameworks? To answer this question
we first specify the hierarchical structure of $M_{\rm q}$ in
terms of the mixing angle $\theta_{\rm q}$ (for q = d or s).
Adopting the radiant
unit for the mixing angles (i.e., $\theta_{\rm u}
\approx 0.085$, $\theta_{\rm d} \approx 0.204$ and
$\theta \approx 0.040$), we have
\begin{eqnarray}
\frac{m_u}{m_c} & \sim & \frac{m_c}{m_t} \;
\sim \; \theta^2_{\rm u} \; , \nonumber \\
\frac{m_d}{m_s} & \sim & \frac{m_s}{m_b} \;
\sim \; \theta^2_{\rm d} \; .
\end{eqnarray}
Then the mass matrices $M_{\rm u}$ and $M_{\rm d}$,
which have the mass scales $m_t$ and $m_b$
respectively, take the following {\it parallel}
hierarchies:
\begin{eqnarray}
M_{\rm u} & \sim & m_t \left ( \matrix{
{\bf 0} & \theta^3_{\rm u} & {\bf 0} \cr
\theta^3_{\rm u} & \theta^2_{\rm u} & \theta^2_{\rm u} \cr
{\bf 0} & \theta^2_{\rm u} & {\bf 1} \cr}
\right ) \; \; , \nonumber \\
\nonumber \\
M_{\rm d} & \sim & m_b \left ( \matrix{
{\bf 0} & \theta^3_{\rm d} & {\bf 0} \cr
\theta^3_{\rm d} & \theta^2_{\rm d} & \theta^2_{\rm d} \cr
{\bf 0} & \theta^2_{\rm d} & {\bf 1} \cr}
\right ) \; \; ,
\end{eqnarray}
where the relevant complex phases have been neglected.
Clearly all three flavor mixing angles can properly
be reproduced from (6.3), once one takes $\theta \approx
\theta^2_{\rm d} \gg \theta^2_{\rm u}$ into account.
The $CP$-violating phase $\varphi$ in $V$ comes
essentially from the phase difference between the
$\theta^3_{\rm u}$ and $\theta^3_{\rm d}$ terms.
Of course $\theta_{\rm u}$ and $\theta_{\rm d}$, which
are more fundamental than the Cabibbo angle $\theta_{\rm C}$
in our point of view, denote perturbative corrections to the
rank-one limits of $M_{\rm u}$ and $M_{\rm d}$
respectively. They are responsible for the generation
of light quark masses as well as the flavor mixing.
They might also be responsible for $CP$ violation in
a specific theoretical framework (e.g., the pure
real $\theta_{\rm u}$ and the pure imaginary $\theta_{\rm d}$
might lead to a phase difference of about $90^{\circ}$
between $M_{\rm u}$ and $M_{\rm d}$, which is just the
source of $CP$ violation favored by current data).
The small parameter $\theta_{\rm q}$ could get its
physical meaning in the Yukawa coupling of
an underlying superstring theory:
$\theta_{\rm q} = \langle \Theta_{\rm q} \rangle /\Omega_{\rm q}$,
where $\langle \Theta_{\rm q} \rangle $ denotes the
vacuum expectation value of the singlet field $\Theta_{\rm q}$,
and $\Omega_{\rm q}$ represents the unification (or string)
mass scale which governs higher dimension operators
(see, e.g., Refs. \cite{Froggatt79,KX98,Ross94}). The quark mass
matrices of the form (6.3) could then be obtained
by introducing an extra (horizontal) U(1) gauge symmetry
or assigning the matter fields appropriately.
A detailed study of possible dynamical models responsible
for the quark mass matrices (5.1) or (6.3) is certainly desirable
but beyond the scope of this work. However, we believe that
the texture zeros and parallel hierarchies of up and down
quark mass matrices do imply specific symmetries, perhaps
at a superhigh scale, and have instructive consequences
on flavor mixing and $CP$-violating phenomena.
The new parametrization of the flavor mixing matrix
that we advocated is particularly useful in studying
the quark mass generation, flavor mixing
and $CP$ violation.
\newpage
\begin{flushleft}
{\Large\bf Appendices}
\end{flushleft}
|
\section{Introduction\label{sec:introduction}}
\nopagebreak
Many spiral galaxies display boxy or peanut-shaped bulges when viewed edge-on.
Unfortunately, statistical studies of the incidence of these objects have not
used objective criteria to quantify the boxiness of the bulges, and they are
therefore hard to compare and yield moderately different results.
Nevertheless, it seems clear that at least 20-30\% of edge-on spirals possess
a boxy/peanut-shaped (hereafter B/PS) bulge (see Jarvis \markcite{j86}1986;
Shaw \markcite{s87}1987; de Souza \& dos Anjos \markcite{dd87}1987). Spiral
galaxies with a B/PS bulge are therefore a significant class of objects.
The fact that the boxy/peanut shape is seen only in edge-on spirals indicates
that the shape is related to the vertical distribution of light. Compared to
the usual $R^{1/4}$ light distribution of spheroids, B/PS bulges have excess
light above the plane at large galactocentric radii (see Shaw
\markcite{s93}1993). Furthermore, the three-dimensional (hereafter 3D) light
distribution of B/PS bulges must be even more extreme than their projected
surface brightness (see, e.g., Binney \& Petrou \markcite{bp85}1985 for
axisymmetric models). B/PS bulges also appear to rotate cylindrically, i.e.\
their rotation is independent of the height above the plane (e.g.\ Kormendy \&
Illingworth \markcite{ki82}1982; Rowley \markcite{r86}1986).
Several models have been proposed to explain the structure of B/PS bulges
(e.g.\ Combes \& Sanders \markcite{cs81}1981; May, van Albada, \& Norman
\markcite{mvn85}1985; Binney \& Petrou \markcite{bp85}1985). However, despite
their prevalence and interesting structural and dynamical properties, B/PS
bulges remain poorly studied observationally, probably because the edge-on
projection makes the interpretation of observational data difficult.
In this paper, we present new kinematical data for a large sample of spiral
galaxies with a B/PS bulge. Our main goals are to determine their 3D structure
and to identify their likely formation mechanism. We use our data to show
that B/PS bulges are simply thick bars viewed edge-on.
In \S~\ref{sec:formation}, we discuss the two main scenarios proposed for the
formation of B/PS bulges -- accretion of satellite galaxies and buckling of a
bar. We describe ways of discriminating between the two scenarios using
kinematical data in \S~\ref{sec:diagnostics}. The observations are presented
in \S~\ref{sec:observations} and the results discussed at length in
\S~\ref{sec:discussion}. We conclude in \S~\ref{sec:conclusions}.
\section{Formation Mechanisms\label{sec:formation}}
\nopagebreak
\subsection{Accretion\label{sec:accretion}}
\nopagebreak
Accretion of external material such as satellite galaxies was the early
favoured mechanism to explain the formation of B/PS bulges in spiral galaxies.
Binney \& Petrou (\markcite{bp85}1985) and Rowley (\markcite{r86}1986,
\markcite{r88}1988) showed that it is possible to construct axisymmetric
cylindrically rotating B/PS bulges from relatively simple distribution
functions. Binney \& Petrou adopted a distribution function with a third
integral of motion, in addition to the energy $E$ and angular momentum along
the symmetry axis $L_z$. This integral favoured orbits reaching a particular
height above the plane. Rowley adopted a two-integral distribution function,
with a truncation depending on both $E$ and $L_z$. Both authors argued that
the required distribution functions can form naturally through the accretion
of material onto a host spiral galaxy. Binney \& Petrou (\markcite{bp85}1985)
additionally argued that, for the accreted material to form a B/PS bulge, the
velocity dispersion of the satellite must be much lower than its orbital
speed, and the decay timescale must be much longer that the orbital time. This
ensures that the accreted material stays clustered in phase space.
Accretion scenarios face several problems. At one extreme, one can consider
the accretion of several small satellite galaxies. However, only a narrow
range of orbital energy and angular momentum can lead to the formation of a
B/PS bulge, so it is improbable that many satellite galaxies would all share
these properties. Remaining satellites should then be present but they are not
observed (Shaw \markcite{s87}1987). The accretion of a single large companion
is also ruled out by the large velocity dispersion and small decay timescale
involved (Binney \& Petrou \markcite{bp85}1985). Similarly, the merger of two
spiral galaxies of similar sizes (or of a spiral and a small elliptical) seems
an unlikely route to form B/PS bulges. This would require a fairly precise
alignment of the spins and orbital angular momenta of the two galaxies. We
recall that about a third of all spiral galaxy bulges should be formed this
way.
From the arguments presented above, it seems that the accretion of a small
number of moderate-sized satellites is the only accretion scenario which may
lead to the formation of B/PS bulges. In favour of accretion is the fact that
the best examples of B/PS bulges are found in small groups (e.g.\ NGC~128,
ESO~597-~G~036). In addition, the possibly related X-shaped galaxies can form
through the accretion of a satellite galaxy (e.g.\ Whitmore \& Bell
\markcite{wb88}1988; Hernquist \& Quinn \markcite{hq89}1989). On the other
hand, no evidence of accretion (arcs, shells, filaments, etc.) were detected
by Shaw (\markcite{s93}1993) near spirals with a B/PS bulge, which argues
against any kind of accretion. Furthermore, spiral galaxies with a B/PS bulge
are not found preferentially in clusters (Shaw \markcite{s87}1987). We are
thus led to the conclusion that, while accretion of external material may play
a role in the formation of some B/PS bulges, it is unlikely to be the primary
formation mechanism.
\subsection{Bar-Buckling\label{sec:buckling}}
\nopagebreak
Bars can form naturally in $N$-body simulations of rotationally supported
stellar disks (e.g.\ Sellwood \markcite{s81}1981; Athanassoula \& Sellwood
\markcite{as86}1986), due to global bisymmetric instabilities (e.g.\ Kalnajs
\markcite{k71}1971, \markcite{k77}1977). Based on 3D $N$-body
simulations, Combes \& Sanders (\markcite{cs81}1981) were the first to suggest
that B/PS bulges may arise from the thickening of bars in the disks of spiral
galaxies. Their results were confirmed and their suggestion supported by later
works (see, e.g., Combes et al.\ \markcite{cdfp90}1990; Raha et al.\
\markcite{rsjk91}1991). In short, the simulations show that, soon after a bar
develops, it buckles and settles with an increased thickness and vertical
velocity dispersion, appearing boxy-shaped when seen end-on and peanut-shaped
when seen side-on. The B/PS bulges so formed are cylindrically rotating, as
required.
Toomre (\markcite{t66}1966) first considered the buckling of disks, in highly
idealised models, and found that, if the vertical velocity dispersion in a
disk is less than about a third of the velocity dispersion in the plane, the
disk will be unstable to buckling modes (fire-hose or buckling instability;
see also Fridman \& Polyachenko \markcite{fp84}1984; Araki
\markcite{a85}1985). Bar formation in a disk makes the orbits within the bar
more eccentric without affecting much their perpendicular motions, thereby
providing a natural mechanism for the bar to buckle. Resonances between the
rotation of the bar and the vertical oscillations of the stars can also make
the disk vertically unstable (see Pfenniger \markcite{p84}1984; Combes et al.\
\markcite{cdfp90}1990; Pfenniger \& Friedli \markcite{pf91}1991). Irrespective
of exactly how a bar buckled, the final shape of the bar is probably due to
orbits trapped around the 2:2:1 periodic orbit family (see, e.g., Pfenniger \&
Friedli \markcite{pf91}1991).
Bar-buckling is the currently favoured mechanism for the formation of B/PS
bulges. In particular, it provides a natural way to form B/PS bulges in
isolated spiral galaxies, which accretion scenarios are unable to do. A number
of facts also suggest a connection between (thick) bars and B/PS bulges,
although they do not support bar-{\em buckling} directly. In particular, the
fraction of edge-on spirals possessing a B/PS bulge is similar to the fraction
of strongly barred spirals among face-on systems ($\approx30\%$; see, e.g.,
Sellwood \& Wilkinson \markcite{sw93}1993; Shaw \markcite{s87}1987). In a few
cases, a bar can also be directly associated with a B/PS bulge from
morphological arguments. NGC~4442 is such an example (Bettoni \& Galletta
\markcite{bg94}1994).
The main observational problem faced by the bar-buckling mechanism is that
B/PS bulges seem to be shorter (relative to the disk diameter) than real bars
or the strong bars formed in $N$-body simulations. However, this might simply
be due to projection (a given surface brightness level being reached at a
smaller radius in a more face-on disk) and no proper statistics have yet been
compiled. The long term secular evolution of bars is also poorly understood.
For example, it is known that bars can transfer angular momentum to a
spheroidal component very efficiently (e.g.\ Sellwood \markcite{s80}1980;
Weinberg \markcite{w85}1985). On the other hand, Debattista \& Sellwood
(\markcite{ds98}1998) showed that, while a bar can be slowed down, it
continues to grow and the boxy/peanut shape is preserved. It is less clear
what happens if a bar is strongly perturbed. However, Norman, Sellwood, \&
Hasan (\markcite{nsh96}1996) showed that, if a bar is destroyed through the
growth of a central mass concentration (e.g.\ Hasan \& Norman
\markcite{hn90}1990; Friedli \& Benz \markcite{fb93}1993), the boxy/peanut
shape is also destroyed.
The merits of the bar-buckling mechanism significantly outweigh these
potential problems, and the bar-buckling scenario to form B/PS bulges remains
largely unchallenged at the moment, despite very little observational support.
The main aim of this paper is to test this mechanism, by looking for bars in a
large sample of spiral galaxies with a B/PS bulge. Although this stops short
of a direct verification of buckling, it does test directly for a possible
relationship between bars and B/PS bulges. We will come back to this
distinction in \S~\ref{sec:discussion}.
Although it is not part of our program, the Milky Way is a primary example of
such a galaxy. The Galactic bulge is boxy-shaped (Weiland et al.\
\markcite{wetal94}1994) and it is now well established that it is bar-like
(e.g.\ Binney et al.\ \markcite{bgsbu91}1991; Paczy\'{n}ski et al.\
\markcite{psuskkmk94}1994; see Kuijken \markcite{k96}1996 for a brief review
of the subject).
We note that ``hybrid'' scenarios have also been proposed to explain the
formation of B/PS bulges, and have gathered recent support from statistical
work on the environment of B/PS bulges by L\"{u}tticke \& Dettmar
(\markcite{ld98}1998). An interaction or merger can excite (or accelerate) the
development of a bar in a disk which is stable (or quasi-stable) against bar
formation (e.g.\ Noguchi \markcite{n87}1987; Gerin, Combes, \& Athanassoula
\markcite{gca90}1990; Miwa \& Noguchi \markcite{mn98}1998). Bars formed this
way are then free to evolve to a boxy/peanut shape in the manner described
above (see, e.g., Mihos et al.\ \markcite{mwhdb95}1995), and the bulges thus
formed owe as much to interactions as they do to the bar-buckling instability.
However, the {\em shape} of the bulges is due to the buckling of the bars and
interaction is merely a way to start the bar formation process. In that sense,
hybrid scenarios for the formation of B/PS bulges are really bar-buckling
scenarios, and the possible accretion of material is not directly related to
the issue of the bulges' shape.
\section{Observational Diagnostics\label{sec:diagnostics}}
\nopagebreak
Our main goal with the observations presented in this paper is to look for the
presence of a bar in the disk of spiral galaxies possessing a B/PS bulge.
There is no straightforward photometric way to identify a bar in an edge-on
spiral. The presence of a plateau in the major-axis light profile of the disk
has often been invoked as an indicator of a bar (e.g.\ de Carvalho \& da Costa
\markcite{dd87}1987; Hamabe \& Wakamatsu \markcite{hw89}1989). However,
axisymmetric or quasi-axisymmetric features (e.g.\ a lens) can be mistaken for
a bar and end-on bars may remain undetected. Kuijken \& Merrifield
(\markcite{km95}1995; see also Merrifield \markcite{mk96}1996) were the first
to demonstrate that bars could be identified kinematically in external edge-on
spiral galaxies. They showed that periodic orbits in a barred galaxy model
produce characteristic double-peaked line-of-sight velocity distributions when
viewed edge-on. This gives their modeled spectra a spectacular
``figure-of-eight'' (or X-shaped) appearance, which they were able to observe
in the long-slit spectra of the B/PS systems NGC~5746 and NGC~5965 (see
Fig.~\ref{fig:main} for examples). Their approach is analogous to that used
in the Galaxy with longitude-velocity diagrams (e.g.\ Peters
\markcite{p75}1975; Mulder \& Liem \markcite{ml86}1986; Binney et al.\
\markcite{bgsbu91}1991).
Bureau \& Athanassoula (\markcite{ba99}1999, hereafter BA99) refined the
dynamical theory of Kuijken \& Merrifield (\markcite{km95}1995). They used
periodic orbit families in a barred galaxy model as building blocks to model
the structure and kinematics of real galaxies. They showed that the global
structure of a position-velocity diagram\footnote{Position-velocity diagrams
(PVDs) show the projected density of material in a system as a function of
line-of-sight velocity and projected position.} (hereafter PVD) taken along
the major axis of an edge-on system is a reliable bar diagnostic, particularly
the presence of gaps between the signatures of the different families of
periodic orbits. Athanassoula \& Bureau (\markcite{ab99}1999, hereafter AB99)
produced similar bar diagnostics using hydrodynamical simulations of the
gaseous component alone. They showed that, when $x_2$ orbits are present
(corresponding to the existence of an inner Lindblad resonance (hereafter
ILR)), the presence of gaps in a PVD, between the signature of the $x_2$-like
flow and that of the outer parts of the disk, reliably indicates the presence
of a bar. If no $x_2$ orbits are present, one must rely on indirect evidence
to argue for the presence of a bar (see, e.g., Contopoulos \& Grosb\o l
\markcite{cg89}1989 for a review of periodic orbits in barred spirals). The
gaps are a direct consequence of the shocks which develop in relatively strong
bars. These shocks drive an inflow of gas toward the center of the galaxies
and deplete the outer (or entire) bar regions (see, e.g., Athanassoula
\markcite{a92}1992). The simulations of \markcite{ab99}AB99 can be directly
compared with the emission line spectra presented here, and will form the
basis of our argument.
The models of \markcite{ba99}BA99 and \markcite{ab99}AB99 can also be used to
determine the viewing angle with respect to a bar, as the signature present in
the PVDs changes with the orientation of the line-of-sight. In particular,
when a bar is seen end-on, the $x_1$ orbits (and $x_1$-like flow, both
elongated parallel to the bar) reach very high radial velocities, while the
$x_2$ orbits (and $x_2$-like flow, both elongated perpendicular to the bar)
show only relatively low velocities. The opposite is true when a bar is seen
side-on. In addition, the presence or absence of $x_2$ orbits can somewhat
constrain the mass distribution and bar pattern speed of an observed galaxy.
We have not developed specific observational criteria to identify past or
current accretion of material in the observed galaxies. As discussed in
\S~\ref{sec:accretion}, accretion will occur through interactions or merger
events. We will take as the signature of such events, and of possible
accretion, the presence of irregularities in the observed PVDs, in particular
strong asymmetries about the center of an object (see Fig.~\ref{fig:main} for
examples).
\section{Observations\label{sec:observations}}
\nopagebreak
\subsection{The Sample\label{sec:sample}}
\nopagebreak
Our sample of galaxies consists of 30 edge-on spirals selected from the
catalogues of Jarvis (\markcite{j86}1986), Shaw (\markcite{s87}1987), and de
Souza \& dos Anjos (\markcite{dd87}1987) (spiral galaxies with a B/PS bulge),
and from the catalogue of Karachentsev, Karachentseva, \& Parnovsky
(\markcite{kkp93}1993) (spirals with extreme axial ratios; see also
Karachentsev et al.\ \markcite{kkpk93}1993). In order to have enough spatial
resolution in the long-slit spectroscopy, but still be able to image the
galaxies relatively quickly with a small-field near-infrared (hereafter NIR)
camera, we have selected objects with bulges larger than 0\farcm6 in diameter
and disks smaller than about 7\farcm0 (at the 25~B~mag~arcsec$^{-2}$ level).
NIR imaging is important to refine the classification of the bulges and to
subsequently study the vertical structure of the identified bars. All objects
are accessible from the south ($\delta\leq15\arcdeg$). Three-quarters (23/30)
of the galaxies either have probable companions or are part of a group or
cluster. A few of these probably are chance alignments, so it is fair to say
that we are not biased either against or for galaxies in a dense environment.
We should therefore be able to estimate reliably the importance of accretion
in the formation of B/PS bulges.
Of the sample galaxies described above, 80\% (24/30) have a boxy or
peanut-shaped bulge, while 20\% (6/30) have a more spheroidal bulge morphology
and constitute a ``control'' sample. Of the former group, it turned out that
17 galaxies have emission lines extending far enough in the disk to apply the
diagnostics developed by \markcite{ba99}BA99 and \markcite{ab99}AB99 with the
ionised gas; all galaxies in the control group fulfill this condition. In this
paper, we will thus concentrate on a main sample of 17 edge-on spiral galaxies
with a B/PS bulge and a comparison sample of 6 edge-on spiral galaxies with
more spheroidal bulges. The galaxies in each sample are listed in
Tables~\ref{ta:main} and \ref{ta:control} respectively, along with information
on their properties and environment. The galaxies with no or confined emission
are listed in Table~\ref{ta:undetected}. For those, stellar kinematics must be
used to search for the presence of bar. We note that the galaxy type listed in
Tables~\ref{ta:main}--\ref{ta:undetected} is not precise to more than one or
two morphological type, because of the difficulty of classifying edge-on
spirals. The bulge-to-disk ratio is effectively the only criterion left to
classify the objects.
Other than the catalogue of Karachentsev et al.\ (\markcite{kkp93}1993), we
are not aware of any general catalogue of edge-on spiral galaxies. This makes
it difficult to build a large and varied control sample including edge-on
spiral galaxies with large bulges (the Karachentsev et al.\
\markcite{kkp93}1993 catalogue is restricted to galaxies with major to minor
axis ratio $a/b\geq7$). Such a catalogue would be very useful, and could
probably be constructed from an initial list of candidates taken from a
catalogue such as the RC3 (de Vaucouleurs et al.\ \markcite{ddcbpf91}1991),
which would then be inspected on survey material.
\subsection{Observations and Data Reduction\label{sec:data}}
\nopagebreak
Our spectroscopic data were acquired between December 1995 and May 1997 (total
of 39~nights) using the Double Beam Spectrograph on the 2.3~m telescope at
Siding Spring Observatory. A $1752\times532$~pixels SITE ST-D06A thinned CCD
was used. The observations discussed in this paper were obtained with the red
arm of the spectrograph centered on the H$\alpha$ $\lambda6563$ emission line.
All galaxies were observed using a $1\farcs8\times400\arcsec$ slit aligned
with the major axis. For objects with a strong dust lane, the slit was
positioned just above it. The spectral resolution is about 1.1~\AA\ FWHM
(0.55~\AA~pixel$^{-1}$) and the spatial scale is 0\farcs9~pixel$^{-1}$. These
data can be directly compared with the gas dynamical models of
\markcite{ab99}AB99.
When no emission line was detected in an object, the red arm of the
spectrograph was moved to the \ion{Ca}{2} absorption line triplet. The blue
arm was always centered on the Mg~$b$ absorption feature. These data will form
the core of a future paper discussing stellar kinematics in the sample
galaxies (including the galaxies in Table~\ref{ta:undetected}). Total exposure
times on both arms ranged from 12000 to 21000~s on each object.
The spectra were reduced using the standard procedure within IRAF. The data
were first bias-subtracted, using both vertical and horizontal overscan
regions, and then using bias frames. If necessary, the data were also
dark-subtracted. The spectra were then flatfielded using flattened continuum
lamp exposures, and wavelength-calibrated using bracketing arc lamp exposures
for each image. The data were then rebinned to a logarithmic wavelength
(linear velocity) scale corresponding to 25~km~s$^{-1}$~pixel$^{-1}$. The
spectra were then corrected for vignetting along the slit using sky exposures.
All exposures of a given object were then registered and offset along the
spatial axis, corrected to a heliocentric rest frame, and combined. The
resulting spectra were then sky-subtracted using source-free regions of the
slit on each side of the objects. The sky subtraction was less than perfect in
some cases, mainly because of difficulties in obtaining a uniform focus of the
spectrograph along the entire length of the slit. This was particularly
troublesome for objects like IC~2531 and NGC~5746 which have sizes comparable
to that of the slit (see, e.g., Fig.~\ref{fig:lineratios}a). In order to
isolate the emission lines, the continuum emission of the objects was then
subtracted using a low-order fit to the data in the spectral direction. The
resulting spectra constitute the basis of our discussion in the next section.
We note that in regions with bright continuum emission, like the center of
some galaxies, the continuum subtraction can leave high shot noise in the
data, which should not be confused with line emission in the grayscale plots
of Figure~\ref{fig:main}--\ref{fig:lineratios}. This is the case for example
in the spectra of ESO~240-~G~011, NGC~1032, and NGC~4703. The effect is
perhaps best seen when a bright star is subtracted, such as in the spectra of
NGC~2788A or NGC~1032 (see Fig.~\ref{fig:main}).
We have not extracted rotation curves from our data. This is because the
entire two-dimensional spectrum, or PVD, is required to identify the signature
of a bar in an edge-on spiral galaxy (see \markcite{ba99}BA99 and
\markcite{ab99}AB99). Evidence of interaction and of possible accretion of
material is also more easily seen in the PVDs. We present the [\ion{N}{2}]
$\lambda6584$ emission line rather than H$\alpha$ because it is not affected
by underlying stellar absorption.
\subsection{Results\label{sec:results}}
\nopagebreak
\placefigure{fig:main}
\placefigure{fig:control}
We present the emission line spectrum for the sample galaxies which have
extended emission only. The PVDs of the galaxies in the main sample and the
control sample are shown in Figure~\ref{fig:main} and
Figure~\ref{fig:control}, respectively. In order to illustrate the range of
galaxy type and bulge morphology in the sample, and to allow connections to be
made between bulge morphologies and kinematical features in the disks, each
PVD is accompanied by a registered image of the corresponding galaxy (from the
Digitized Sky Survey) on the same spatial scale. We discuss here the trends
observed across the data set.
The most important trend, and the main result of this paper, is that most
galaxies in the main sample show a clear bar signature in their PVD (as
described in \S~\ref{sec:diagnostics}). Of the 17 galaxies in the main sample
of spirals with a B/PS bulge and extended emission lines, we conclude that 14
are barred. In these objects, the PVD clearly shows a strong and steep inner
component, associated with an $x_2$-like flow, and a slowly-rising almost
solid-body component, associated with the outer disk, and joining the flat
section of the rotation curve in the outer parts. The two components are
separated by a gap, caused by the absence (or low density) of gaseous material
with $x_1$-like kinematics in the outer bar regions\footnote{The PVDs of
NGC~128 and IC~2531 do not display a bar signature, but Emsellem \&
Arsenault (\markcite{ea97}1997) and Bureau \& Freeman (\markcite{bf97}1997)
showed, using other data, that each galaxy harbours a bar.}. The best
examples of this type of bar signature are seen in the PVD of earlier-type
objects, like NGC~2788A, NGC~5746, and IC~5096. However, the signature is
still clearly visible in the PVD of galaxies as late as ESO~240-~G~011.
In the main sample, only one galaxy, NGC~4469, may be axisymmetric, with no
evidence of either a bar or interaction (although it is not possible to rule
out an interaction which would have occured a long time ago, leaving no
observable trace). Two galaxies, NGC~3390 and ESO~597-~G~036, have a disturbed
strongly asymmetric PVD, which we ascribe to a recent interaction (obvious in
the case of ESO~597-~G~036). These interactions may have led to the accretion
of material. The results for the entire main sample are summarised in
Table~\ref{ta:main}.
Another significant result of this study is that no galaxy in the control
sample shows evidence for a bar. Four of the six galaxies appear to be
axisymmetric, without evidence for either a bar or interaction, and two
galaxies (NGC~5084 and NGC~7123) have a disturbed PVD, indicating they
underwent an interaction recently and possibly accreted material. These
results are tabulated in Table~\ref{ta:control}.
\section{Discussion\label{sec:discussion}}
\nopagebreak
\subsection{The Structure of Boxy/Peanut-Shaped Bulges\label{sec:structure}}
\nopagebreak
The only previous study of this kind was that of Kuijken \& Merrifield
(\markcite{km95}1995), who proposed the method and considered two galaxies;
Bureau \& Freeman (\markcite{bf97}1997) presented preliminary results of the
current work. This is thus the first systematic observational study of the
relationship between bars and B/PS bulges. In summary, our main result is
that, based on new kinematical data, 14 of the 17 galaxies with a B/PS bulge
in our sample are barred, and the remaining 3 galaxies show evidence of
interaction or may be axisymmetric. None of the 6 galaxies without a B/PS
bulge in our sample shows any indication of a bar. This means that most B/PS
bulges are due to the presence of a thick bar viewed edge-on, and only a few
may be due to the accretion of external material. In addition, the more
spheroidal bulges (i.e.\ non-B/PS) do seem axisymmetric. It appears then that
most B/PS bulges are edge-on bars, and that most bars are B/PS when viewed
edge-on. However, the small size of the control sample prevents us from making
a stronger statement about this converse. There is also a continuum of bar
strengths in nature and we would expect to have intermediate cases. The
galaxies NGC~3957 and NGC~4703 may represent such objects: one could argue
that their PVDs display weak bar signatures, and indeed their bulges are the
most flattened in the control sample.
If bars were unrelated to the structure of bulges, we would have expected only
about 5 galaxies in the main sample to be strongly barred, and about 2
galaxies in the control sample (about 30\% of spirals are strongly barred,
Sellwood \& Wilkinson \markcite{sw93}1993). Clearly, our results are
incompatible with these expectations. Recent results by Merrifield \& Kuijken
\markcite{mk99}(1999) also support our conclusions. Based on a smaller sample
of northern edge-on spirals, they show clearly that as bulges become more
B/PS, the complexity and strength of the bar signature in the PVD also
increase.
Our association of bars and B/PS bulges supports the bar-buckling mechanism
for the formation of the latter. However, we do not test directly for
buckling, but rather for the presence of a barred potential in the plane of
the disks. Because bars form rapidly and buckle soon after, on a timescale of
only a few dynamical times (see, e.g., Combes et al.\ \markcite{cdfp90}1990),
it is unlikely that any galaxy in this nearby sample would have been caught in
the act. Thus, other mechanisms which could lead to thick bars cannot be
excluded. In addition, we have no way of determining how the bars themselves
formed, or even whether they formed spontaneously in isolation or through
interaction with nearby galaxies or companions. Therefore, the possibility of
hybrid scenarios for the formation of B/PS bulges, where a bar is formed
because of an interaction and subsequently thickens due to the buckling
instability, remains (see \S~\ref{sec:buckling}).
We have looked mainly at objects with large bulges; only two of the galaxies
studied are late-type spirals (ESO~240-~G~011 and IC~5176). This is a
selection effect due to the difficulty to identify very small B/PS bulges. It
would therefore be interesting to search for bars in very late edge-on spiral
galaxies, and verify whether thin bars do exist: the bar in ESO~240-~G~011 is
not very thick, but it is thicker (isophotally) than the disk. \ion{H}{1}
synthesis imaging is probably the best way to achieve this goal, as these
objects are often dusty and \ion{H}{1}-rich. Such bars may even provide a
novel way to constrain the total (luminous and dark) mass distribution of
spirals, in a manner analogous to the use of warps or flaring, as buckling is
sensitive to the presence of a dark halo (e.g.\ Combes \& Sanders
\markcite{cs81}1981). We will report on \ion{H}{1} synthesis observations of
a few objects in our sample in a later paper.
Our observations revive the old issue of the exact nature of bulges. In
face-on systems, one can often clearly identify a bar and a more nearly
axisymmetric component usually referred to as the bulge. However, Kormendy
(\markcite{k93}1993) has argued that, at least in some examples, these
apparent bulges may just be structures in the disk. In edge-on spirals, we are
not aware of any galaxies displaying two separate vertically extended
components. This raises the question of whether the bars and bulges of
face-on systems are really two distinct structural and dynamical components,
despite the fact that they can be separated photometrically. Our data on
edge-on galaxies tightly link the presence of a bar with the presence of a
B/PS bulge, which suggests that bars and B/PS bulges are very closely related.
This view is supported by theoretical and modelling work on barred spiral
galaxies (e.g.\ Pfenniger \markcite{p84}1984; Pfenniger \& Friedli
\markcite{pf91}1991), as well as by some photometric studies (e.g.\ Ohta
\markcite{o96}1996). However, more work is required to settle the issue.
Kinematical data covering whole bulges would be particularly useful.
\markcite{ba99}BA99 and \markcite{ab99}AB99 proposed diagnostics, again based
on the structure of the observed PVD, to determine the viewing angle with
respect to a bar in an edge-on disk (see \S~\ref{sec:diagnostics}). When the
bar is seen close to side-on, the maximum line-of-sight velocity reached by
the $x_2$-like flow is similar to or larger than the flat portion of the
rotation curve, and the component of the PVD associated with that flow is very
steep. When the bar is seen close to end-on, the $x_2$-like flow only reaches
low velocities and extends to relatively large projected distances. These
diagnostics are ideally suited to test the main prediction of $N$-body models,
that bars are peanut-shaped when seen side-on and boxy-shaped when seen end-on
(see, e.g., Combes \& Sanders \markcite{cs81}1981; Raha et al.\
\markcite{rsjk91}1991).
Of the 12 barred galaxies in the main sample for which it is possible to apply
these criteria (we exclude NGC~128 and IC~2531), two-thirds (8/12) seem to be
consistent with the above prediction of $N$-body simulations. For example, in
the galaxy with a peanut-shaped bulge IC~4937, it is clear that the steep
inner component associated with the $x_2$-like flow extends to higher
velocities than the outer parts of the disk (see Fig.~\ref{fig:main}). This
situation is reversed in NGC~1886, which has a boxy-shaped bulge. However,
caution is required when interpreting this result. Firstly, the present
classification of the shape of the bulges is affected by both dust and the low
dynamic range of the material used (Digitized Sky Survey). To remedy this
problem, we have acquired $K$-band images of all the sample galaxies and will
report on these observations in a future paper. Secondly, no clear prediction
has been made from $N$-body simulations about the viewing angle at which the
transition from a boxy to a peanut-shaped bulge occurs. In that regard, it
would be useful to apply quantitative measurements of the boxiness and
``peanutness'' of the bulges to both simulation results and observational data
(see, e.g., Bender \& M\"{o}llenhoff \markcite{bm87}1987 and Athanassoula et
al.\ \markcite{amwpplb90}1990 for two possible methods). Thirdly, because the
$x_2$-like flow is located near the center of the galaxies, the velocities it
reaches depend somewhat on the central concentration of the objects (which
affects the circular velocity in the central regions). This obviously varies
significantly amongst the galaxies in our sample. Therefore, while our
observations may support the prediction of $N$-body models concerning the
orientation of the bar in B/PS bulges, we believe that it is premature to make
a detailed quantitative comparison of the data with the models.
For a more detailed comparison with theory, data of higher signal-to-noise and
higher spatial resolution than the average PVD presented here would be very
desirable. However, it would be worthwhile to model individually the best PVDs
obtained in the present study (e.g.\ NGC~5746, IC~5096, and a few others).
This would very likely lead to tight constraints on the mass distribution and
bar properties of the galaxies, including the orientation and pattern speed of
the bars (see \markcite{ba99}BA99; \markcite{ab99}AB99). The $K$-band images
could also be used to constrain the mass distributions. Comparing the data
with the kinematics (or simply the rotation curve) predicted from an
axisymmetric deprojection would provide an easy test of the shape of the
bulges. On a related subject, the significant thickness of bars suggests that
their 3D structure should be taken into account when deriving the potential of
more face-on systems from NIR images.
In that regard, we should stress that the bar diagnostics we used rely on the
presence of an $x_2$-like flow in the center of the galaxies, and thus on the
existence of ILRs (or, at least, one ILR; see \markcite{ba99}BA99;
\markcite{ab99}AB99). A priori, barred disks or B/PS bulges need not have
ILRs, but at least 13 of the 17 galaxies in the main sample do (we
additionally exclude NGC~128 here). Our data therefore strongly support the
view that barred spiral galaxies generally have ILRs (see also Athanassoula
\markcite{a91}1991, \markcite{a92}1992).
\subsection{Dust and Emission Line Ratios in Boxy/Peanut-Shaped
Bulges\label{sec:elratios}}
\nopagebreak
Because many galaxies in our sample have a prominent dust lane, it is
important to consider the effects dust may have on our observations. Its
principal consequence in edge-on systems is to limit the depth to which the
line-of-sight penetrates the disk. To bypass this problem, we selected many
galaxies to be slightly inclined, so it was possible in those objects to
position the slit just above the dust lane and have a line-of-sight that still
goes through most of the disk, as required for a comparison with the models of
\markcite{ba99}BA99 and \markcite{ab99}AB99. In the few cases with a strong
dust lane and a perfectly edge-on disk, we tried again to offset the slit
slightly. Unfortunately, with the observational set-up available at the
telescope, it was difficult to position the slit with great precision. The
objects where we suspected that dust could affect our observations are
indicated in Tables~\ref{ta:main}--\ref{ta:undetected}. A large dust optical
depth produces an almost featureless PVD, as one sees only an outer annulus of
material in the disk, and the rotation curve appears slowly-rising and
solid-body (see, e.g., Bosma et al.\ \markcite{bbfa92}1992). The only objects
showing obvious signs of extinction are IC~2531\footnote{This is confirmed by
the \ion{H}{1} radio synthesis data of Bureau \& Freeman
(\markcite{bf97}1997), which reveal a complex PVD with a bar signature.},
NGC~4703, and possibly ESO~443-~G~042. Because we see a lot of structure in
most PVDs, including the PVDs of objects with a significant dust lane, we do
not believe that our results are significantly affected by dust. We do detect
a clear bar signature in most galaxies in the main sample.
This statement is strengthened by the fact that all the PVDs showing a bar
signature are close to symmetric. \markcite{ab99}AB99 showed that the bar
signature would be strongly asymmetric in a very dusty disk, and this is not
observed. Similarly, it is improbable that irregular dust distributions would
lead to such well-ordered and symmetric PVDs (very large and localized dust
``patches'' would be required to create the important gaps observed in many
objects).
\placefigure{fig:lineratios}
An unexpected but interesting prospect raised by our observations concerns
emission line ratios. For many of the barred galaxies in the main sample, the
emission line ratios in the central regions are different from those expected
of typical \ion{H}{2} regions. For 9 barred galaxies out of 14, mostly those
with the strongest bar signatures, the [\ion{N}{2}] $\lambda6584$ to H$\alpha$
$\lambda6563$ ratio in the bulge region is greater than unity. In particular,
in a few objects, the steep inner component of the PVD, associated with the
$x_2$-like flow, is much stronger in [\ion{N}{2}] than it is in H$\alpha$,
while the slowly rising component, associated with the outer disk, has a
[\ion{N}{2}]/H$\alpha$ ratio typical of \ion{H}{2} regions. In fact, the inner
component can be almost absent in H$\alpha$. We illustrate these behaviours in
Figure~\ref{fig:lineratios}, which shows the PVDs of the galaxies NGC~5746 and
IC~5096 in the H$\alpha$ and [\ion{N}{2}] $\lambda6548,6584$ lines.
It is possible that the H$\alpha$ emission line is weakened by the underlying
stellar absorption. This suggestion is supported by the fact that the spectra
of many of the galaxies in the main sample show strong Balmer absorption
lines. However, the H$\alpha$ absorption would need to be very large to
account for the extreme [\ion{N}{2}]/H$\alpha$ ratios observed in some objects
(e.g.\ IC~5096). The strong Balmer lines observed in many objects are
nevertheless interesting in themselves, and indicate that a significant
intermediate age ($\sim1$~Gyr) stellar population is present in the central
regions of the disk of many of the galaxies. It would be interesting to
investigate if these past bursts of star formation can be related to the
presence of the bars.
The high emission line ratios are interesting for two reasons. Firstly, high
[\ion{N}{2}]/H$\alpha$ ratios are commonly believed to be produced by shocks
(see, e.g., Binette, Dopita, \& Tuohy \markcite{bdt85}1985; Dopita \&
Sutherland \markcite{ds96}1996). This is consistent with the view that B/PS
bulges are barred spirals viewed edge-on. The steep inner components of the
PVDs, which display high [\ion{N}{2}]/H$\alpha$ ratios, are associated with an
$x_2$-like flow and the nuclear spirals observed in many barred spiral
galaxies (\markcite{ab99}AB99). Athanassoula (\markcite{a92}1992) showed
convincingly that these are the locus of shocks. Secondly, if one were to
derive H$\alpha$ and [\ion{N}{2}] rotation curves from the data, by taking the
upper envelope of the PVDs (the standard method), the H$\alpha$ and [\ion{N}{2}]
rotation curves would significantly differ for many objects. The [\ion{N}{2}]
lines would yield rapidly rising rotation curves flattening out at small
radii, while the H$\alpha$ line would yield slowly rising rotation curves
flattening out at relatively large radii. Mass models derived from such data
would thus yield qualitatively different results, and our understanding of
galactic dynamics and structure gained from this type of work could be
seriously erroneous (at least for highly inclined spirals). Of course, now
that these galaxies are known to be barred, their rotation curves should not
be used directly for mass modelling, as they are not a good representation of
the circular velocity.
Data such as those presented in Figure~\ref{fig:lineratios} also open up the
possibility of determining the ionisation conditions and abundance of the gas
in different regions of the galaxies in a single observation. Because the
deprojected location of each component of the PVDs is known (see
\markcite{ab99}AB99), this provides an efficient way to study the effects of
bars on the interstellar medium of galaxies on various scales.
\section{Conclusions\label{sec:conclusions}}
\nopagebreak
In this paper, we discussed the various mechanisms proposed for the formation
of the boxy and peanut-shaped bulges observed in some edge-on spiral galaxies.
We argued that accretion scenarios were unlikely to account for most
boxy/peanut-shaped (B/PS) bulges, but that bar-buckling scenarios, discovered
through $N$-body simulations, had this potential. Using recently developed
kinematical bar diagnostics, we searched for bars in a large sample of edge-on
spiral galaxies with a B/PS bulge. Of the 17 galaxies where the diagnostics
could be applied using emission lines, 14 galaxies were shown to be barred, 2
galaxies were significantly disturbed, and 1 galaxy seemed to be axisymmetric.
In a control sample of 6 galaxies with spheroidal bulges, none appeared to be
barred.
Our study supports the bar-buckling mechanism for the formation of B/PS
bulges. Our results imply that most B/PS bulges are due to the presence of a
thick bar that we are viewing edge-on, while only a few may be due to the
accretion of external material. Furthermore, spheroidal bulges do appear to be
axisymmetric. This suggests that all bars are B/PS. Our observations also seem
to support the main prediction of $N$-body simulations, that bars appear
boxy-shaped when seen end-on and peanut-shaped when seen side-on. However,
this issue should be revisited in a more quantitative manner in the future.
With our data, we have no way of determining whether the bars leading to B/PS
bulges have formed in isolation or through interactions and mergers. The
association of B/PS bulges and bars is entirely consistent with the properties
of the bulge of the Milky Way, which is known to be both boxy and bar-like.
We considered the effects of dust on our observations, but concluded that it
does not affect our results significantly. We have also shown that emission
line ratios correlate with kinematical structures in many barred galaxies.
This make possible a direct study of the large scale effects of bars on the
interstellar medium in disks.
Our study opens up the possibility to study observationally the vertical
structure of bars. This was not possible before and represents an interesting
spin-off from the use of bar diagnostics in edge-on spiral galaxies. To this
end, we have obtained $K$-band images of all our sample galaxies, and will
report on this work in a future paper.
\acknowledgments
We thank the staff of Mount Stromlo and Siding Spring Observatories for their
assistance during and after the observations. We also thank A.\ Kalnajs, E.\
Athanassoula, A.\ Bosma, and L.\ Sparke for useful discussions. M.\ B.\
acknowledges the support of an Australian DEET Overseas Postgraduate Research
Scholarship and a Canadian NSERC Postgraduate Scholarship during the conduct
of this research. The Digitized Sky Surveys were produced at the Space
Telescope Science Institute under U.S. Government grant NAG W-2166. The images
of these surveys are based on photographic data obtained using the Oschin
Schmidt Telescope on Palomar Mountain and the UK Schmidt Telescope. The plates
were processed into the present compressed digital form with the permission of
these institutions. The NASA/IPAC Extragalactic Database (NED) is operated by
the Jet Propulsion Laboratory, California Institute of Technology, under
contract with the National Aeronautics and Space Administration.
|
\section{Introduction}
Many strong interaction--processes involve meson--baryon coupling constants
as the main ingredient. The determination of these fundamental quantities
requires information about the physics at large distance. In other words,
for a reliable determination of these parameters we need some
nonperturbative approach. Among all nonperturbative approaches, QCD sum
rules \cite{R1} is one of the most powerful method in studying
the properties of hadrons. This
method is based on the short distance OPE of vacuum--vacuum correlation
function in terms of condensates. For the processes involving light mesons
$\pi,~K$ or $\rho$, there is an alternative method to the traditional QCD
sum rules, namely, light cone QCD sum rules \cite{R2}. In this approach the
expansion of the vacuum--meson correlator is performed near the light cone
in terms of the meson wave functions. The meson wave functions are defined
by the matrix elements of non--local composite operators sandwiched between
the meson and vacuum states and classified by their twists, rather
than dimensions of the operators, as is the case in the traditional sum
rules. Many applications of light--cone QCD sum rules can be found in
\cite{R3}--\cite{R11} and references therein.
In this work we use light cone QCD sum rules approach for determination
of the coupling constants of the pion to the lowest states of the baryon
octet $\Sigma$ and $\Lambda$, $g_{\pi\Sigma\Lambda}$ and
$g_{\pi\Sigma\Sigma}$. Note that these coupling constants were investigated
in framework of the QCD sum rules based on pion--to--vacuum matrix element
in the leading order of the pion momentum $q$ for the structure
$\not\!q \gamma_5$ in \cite{R12}, where $q$ is the pion momentum.
The results of this work are currently under debate in literature (see
discussions in \cite{R13} and \cite{R14}).
Moreover in \cite{R15} and \cite{R16}
it was pointed out that there is
coupling scheme dependence for the structures $\gamma_5$, $\not\!q \gamma_5$,
i.e., dependence on the pseudoscalar or pseudovector forms of the effective
interaction Lagrangian of pion with hadrons
have been used, while the structure $\sigma_{\mu\nu} \gamma_5$ is shown to
be independent of any coupling schemes. For this reason, in present work we
choose the structure $\sigma_{\mu\nu} \gamma_5 p^\mu q^\nu$, where $p$ and
$q$ are the $\Lambda~(\Sigma)$ and the pion momenta, respectively. It should
be noted that the sum rules for the $\sigma_{\mu\nu} \gamma_5 p^\mu q^\nu$
structure was derived in \cite{R17} in investigation of $g_{\rho\omega\pi}$
coupling constant.
The paper is organized as follows. In Section 2 we derive sum rules for the
pion--baryon coupling constants $g_{\pi\Sigma\Lambda}$ and
$g_{\pi\Sigma\Sigma}$ for the structure
$\sigma_{\mu\nu} \gamma_5 p^\mu q^\nu$. Section 3 is devoted to the
numerical analysis of the sum rules for $g_{\pi\Sigma\Lambda}$ and
$g_{\pi\Sigma\Sigma}$ and discussion.
\section{Formulation of the pion--baryon sum rule for the
$g_{\pi\Sigma\Lambda}$ and $g_{\pi\Sigma\Sigma}$}
According to the main philosophy of the QCD sum rules, a
quantitative estimation for $g_{\pi\Sigma\Lambda}$ and
$g_{\pi\Sigma\Sigma}$ couplings can be obtained by
matching the representations of a suitable correlator in terms of hadronic
(physical part) and quark--gluon language (theoretical part). For this
purpose we consider the following two--point correlator function with pion
\begin{eqnarray}
\Pi(p,q) = \int d^4x \, e^{ipx} \left< \pi(q) \left| \mbox{\rm T}
\left[\eta_Y (x) \bar \eta_{\Sigma^+} (0) \right] \right| 0 \right>~,
\end{eqnarray}
where $p$ and $\eta_Y$ are the four--momentum of the hyperon (in our case
$\Lambda^0$ or $\Sigma^0$) and its interpolation current, respectively,
$\eta_{\Sigma^+}$ is the interpolating current of $\Sigma^+$ and
$q$ is the pion four--momentum.
The interpolating currents for $\Lambda^0$, $\Sigma^0$ and $\Sigma^+$
are \cite{R18}
\begin{eqnarray}
\eta_{Y^0} &=& \alpha \, \epsilon_{abc} \left[
\left( u_a^T {\cal C} \gamma_\mu s_b \right) \gamma_5 \gamma^\mu d_c \mp
\left( d_a^T {\cal C} \gamma_\mu s_b \right) \gamma_5 \gamma^\mu u_c \right]~,\nonumber \\
\eta_{\Sigma^+}&=& \epsilon_{abc}
\left( u_a^T {\cal C} \gamma_\mu u_b \right) \gamma_5 \gamma^\mu s_c~,
\end{eqnarray}
where $s,~u$ and $d$ are strange, up and down quark fields,
the upper(lower) sign corresponds to $\Lambda^0(\Sigma^0)$ and
$\alpha=\sqrt{2/3}$ for $\Lambda^0$ and $\sqrt{2}$ for $\Sigma^0$,
respectively, $a,~b,~c$ are the color indices, ${\cal C}$ is the charge conjugation
operator.
Saturating correlator (1) with the $Y~(=\Lambda^0$ or $\Sigma^0$)
and $\Sigma^+$ states in the phenomenological part, we get
\begin{eqnarray}
\Pi=
\frac{\left< \pi(q) Y \vert \Sigma^+ \right> \,
\left< \Sigma^+(p+q) \vert \bar \eta_{\Sigma^+} \vert 0 \right>
\, \left< 0\vert \bar \eta_Y \vert Y(p) \right>}
{(p^2-m_Y^2) \left[(p+q)^2 - m_{\Sigma^+}^2\right]} +
\mbox{\rm high. reson.}
\end{eqnarray}
The matrix elements in Eq. (3) are defined in the following way
\begin{eqnarray}
\left< 0 \left| \eta_Y(x) \right| Y(p) \right> &=& \lambda_Y u(p)~, \nonumber \\
\left< \Sigma^+(p+q) \left| \bar \eta_{\Sigma^+} \right| 0 \right> &=&
\lambda_\Sigma \bar u(p+q)~,\nonumber \\
\left< \pi(q) Y(p)\vert \Sigma^+(p+q) \right> &=& -
g_{Y\Sigma^+\pi^-} \, \bar u(p) \gamma_5 u(p+q)~.
\end{eqnarray}
Substituting Eq. (4) in Eq. (3), and choosing the structure
$i\, \sigma_{\alpha\beta} p^\alpha q^\beta \gamma_5$
for the physical part of Eq. (1), we get
\begin{eqnarray}
\Pi^{phys}= - \frac{g_{\pi^-Y\Sigma^+} \lambda_Y \lambda_{\Sigma^+}}
{(p^2-m_Y^2) \left[(p+q)^2 - m_\Sigma^2\right]}
+ \mbox{\rm higher resonances}~.
\end{eqnarray}
Let us now consider the theoretical part of the correlator (1). From
this correlator we have (we present only the terms which give contributions
to the above--mentioned Lorentz structure)
\begin{eqnarray}
\Pi &=& - \alpha \int d^4x\,e^{ipx} \Big\{ - \gamma_5 \gamma_\mu
\gamma_5 \gamma_\varphi \gamma_\rho\,
{\cal C}{\cal S}^T {\cal C}^{-1} \gamma_\mu {\cal S}^s \gamma_\rho
\gamma_5 \left< \pi \left| \bar u \gamma_5 \gamma_\varphi d \right| 0 \right> \nonumber \\
&&+\frac{1}{2} \gamma_5 \gamma_\mu \sigma_{\alpha\beta} \gamma_\rho\,
{\cal C}{\cal S}^T {\cal C}^{-1} \gamma_\mu {\cal S}^s \gamma_\rho
\gamma_5 \left< \pi \left| \bar u \sigma_{\alpha\beta} d \right| 0 \right> \nonumber \\
&& \pm \Big[ - \gamma_5 \gamma_\mu {\cal S} \gamma_\rho \,{\cal C}
(\gamma_5 \gamma_\varphi)^T
{\cal C}^{-1} \gamma_\mu {\cal S}^s \gamma_\rho \gamma_5
\left< \pi \left| \bar u \gamma_5 \gamma_\varphi d \right| 0 \right> \nonumber \\
&&+ \frac{1}{2} \gamma_5 \gamma_\mu {\cal S} \gamma_\rho \,{\cal C}
\sigma_{\alpha\beta}^T {\cal C}^{-1} \gamma_\mu {\cal S}^c \gamma_\rho \gamma_5
\left< \pi \left| \bar u \sigma_{\alpha\beta}d \right| 0\right>\Big] \Big\} ~,
\end{eqnarray}
where upper(lower) sign corresponds to $\Lambda(\Sigma)$ case and
$\alpha=\sqrt{2/3}$ for $\Lambda$ while $\alpha=\sqrt{2}$ for $\Sigma$.
Here ${\cal S}$ and ${\cal S}^s$ are the propagators containing both
perturbative and nonperturbative contributions, respectively. Here we
present the explicit form of $i {\cal S}^s(x)$
\begin{eqnarray}
\lefteqn{
i {\cal S}^s(x) = \frac{i}{2 \pi^2} \frac{\not\!x}{x^4} -
\frac{m_s}{4\pi^2} \frac{1}{x^2} - \frac{1}{12}
\langle \bar s s \rangle \left( 1 - \frac{i m_s}{4} \not\!x \right)} \nonumber \\
&& - \frac{1}{192} m_0^2 \langle \bar s s \rangle
\left( 1 - \frac{i m_s}{6} \not\!x \right) -
i g_s \frac{1}{16 \pi^2} \int_0^1 du \left\{
\frac{\not\!x}{x^2} \sigma_{\alpha\beta} G^{\alpha\beta}(ux) -
4i \frac{x_\alpha}{x^2}G^{\alpha\beta} \gamma_\beta \right\}~,
\end{eqnarray}
where $m_s$ is the mass of the strange quark and $G_{\alpha\beta}$
is the gluon field strength tensor.
The form of ${\cal S}$ can be obtained from Eq. (7) by making the
replacements $\langle \bar s s \rangle \rightarrow \langle \bar q q \rangle$ and
$m_s \rightarrow 0$. From Eq. (6) we observe that, in calculation of the
correlator function in QCD, the matrix element of the nonlocal operators
between the vacuum and pion states are needed. These matrix elements define
the two particle pion wave functions and up to twist four they can be
written as \cite{R6,R7}
\begin{eqnarray}
\left< \pi(q) \left| \bar d \gamma_\mu \gamma_5 u \right| 0 \right> &=&
- i f_\pi q_\mu \int_0^1 du \, e^{iuqx} \big[ \varphi_\pi(u) + x^2 g_1(u) \big]
\nonumber \\
&&+f_\pi \left( x_\mu - \frac{x^2 q_\mu}{qx} \right)
\int_0^1 du\, e^{iuqx} g_2(u)~, \nonumber \\ \nonumber \\
\left< \pi(q) \left| \bar d (x) \sigma_{\alpha\beta}\gamma_5 u(0) \right| 0 \right> &=&
i \left( q_\alpha x_\beta - q_\beta x_\alpha \right)
\frac{f_\pi m_\pi^2}{6(m_u+m_d)}
\int_0^1 du \,e^{iuqx} \varphi_\sigma(u)~.
\end{eqnarray}
Using Eqs. (6), (7) and (8) we get the following result
for theoretical part (for the
structure $i \sigma_{\alpha\beta} x_\alpha q^\beta \gamma_5$)
\begin{eqnarray}
\lefteqn{
\Pi^{theor} =} \nonumber \\
&& - \alpha f_\pi \int_0^1 dx \,e^{ipx} \Bigg\{
\Bigg[ (1 \pm 1) \frac{m_s}{2 \pi^4 x^6} +
(-\lambda \pm \sigma) \left( \frac{1}{6 \pi^2 x^4} +
\frac{m_0^2}{96 \pi^2 x^2} \right) \Bigg]
\int_0^1 du\, e^{iuqx} \varphi_\pi(u) \nonumber \\
&&+\Bigg[ (1 \mp 1) \frac{\mu_\pi}{6 \pi^4 x^6}
+ (1 \pm 1) \frac{\mu_\pi \left< g^2 G^2 \right>}{2304 \pi^4 x^2}+
(1 \mp 1) \frac{m_s \mu_\pi \left< \bar s s \right>}{144 \pi^2 x^2}
\pm \frac{m_s \mu_\pi \sigma}{72 \pi^2 x^2}
\Bigg] \int_0^1 du\, e^{iuqx} \varphi_\sigma(u) \nonumber \\
&& +\Bigg[ \frac{1}{6 \pi^2 x^2} (-\lambda \pm \sigma)
+ (1 \pm 1) \frac{m_s}{2 \pi^4 x^4} \Bigg]
\int_0^1 du\, e^{iuqx} \left[ g_1(u) + G_2 (u) \right] \Bigg\}~,
\end{eqnarray}
where
\begin{eqnarray}
G_2(u) &=& - \int_0^u g_2(v) dv~, \nonumber \\
\lambda &=& \langle \bar q q \rangle - \langle \bar s s \rangle~, \nonumber \\
\sigma &=& \langle \bar q q \rangle + \langle \bar s s \rangle~. \nonumber
\end{eqnarray}
Our next task is to perform integration over $x$ and perform double Borel
transformation in Eq. (9) with respect to the variables $p^2$ and $(p+q)^2$,
in order to get an answer for the theoretical part of the sum rules. As an
example, let us demonstrate on one of the terms in Eq. (9) how integration
over $x$ and double Borel transformation can be carried. Consider the
following term
\begin{eqnarray}
\int du \varphi_\pi(u) \int d^4 x \frac{e^{i(p+qu)x}}{x^2} i
\sigma_{\alpha\beta} x_\alpha q_\beta~. \nonumber
\end{eqnarray}
In performing the $x$ integration, we will make use of the
formula \cite{R19}
\begin{eqnarray}
\int \frac{d^4 x}{(x^2)^n} e^{ipx} =
\frac{i (-1)^n 2^{4-2n} \pi^2}{\Gamma(n-1) \Gamma(n)}
(p^2)^{n-2} ln(-p^2) + {\cal P}_{n-2}~,~~~(n \ge 2 )~, \nonumber
\end{eqnarray}
where ${\cal P}_{n-2}$ is a polynomial of power $n-2$. However this polynomial
is inessential, since it vanishes after double Borel transformation. Hence,
disregarding this polynomial we have
\begin{eqnarray}
\int du \varphi_\pi(u) \int d^4 x \frac{e^{i(p+qu)x}}{x^2}
i \sigma_{\alpha\beta} x_\alpha q_\beta &=&
\int du \varphi_\pi(u) \left( i \frac{\partial}{\partial p_\alpha}
\frac{\partial}{\partial p_\rho} \frac{\partial}{\partial p_\rho} \right)
\int d^4 x \frac{e^{i(p+qu)x}}{x^4} i \sigma_{\alpha\beta} q_\beta \nonumber \\
&=& 8 \int du \varphi_\pi(u)\frac{P_\alpha}{P^4} i \sigma_{\alpha\beta} q_\beta ~, \nonumber
\end{eqnarray}
where $P=p + qu$. The last step in this calculation is performing double
Borel transformation over the variables $p^2$ and $(p+q)^2$ to the
expression
\begin{eqnarray}
\int du \varphi_\pi(u) \frac{1}{\left[ \left( p+qu \right)^2 \right]^2}~. \nonumber
\end{eqnarray}
Rewriting $(p+qu)^2 = p^2 \bar u + u (p+q)^2 $ (here the pion mass is
neglected) and using the exponential representation for the denominator,
\begin{eqnarray}
\frac{1}{\left[ p^2 \bar u + u (p+q)^2 \right]^2} =
\int_0^\infty d \alpha \alpha e^{-\alpha \left[ p^2 \bar u +
u (p+q)^2 \right]}~, \nonumber
\end{eqnarray}
we have
\begin{eqnarray}
\int du \varphi_\pi(u) \frac{1}{\left[ \left( p+qu \right)^2 \right]^2} =
\int du \varphi_\pi(u) \int d \alpha \alpha e^{-\alpha \left[ p^2 \bar u +
u (p+q)^2 \right]}~. \nonumber
\end{eqnarray}
The double Borel transformation over the variable $p^2$ and $(p+q)^2$ is
done with the help of the following general formula
\begin{eqnarray}
{\cal B}_{p^2}^{M_1^2} {\cal B}_{(p+q)^2}^{M_2^2}
\frac{\Gamma(n)}{\left[ - \bar u p^2 - (p+q)^2 u \right]^n} =
(M^2)^{2-n} \delta (u-u_0)~,\nonumber
\end{eqnarray}
where
\begin{eqnarray}
M^2 = \frac{M_1^2 M_2^2}{M_1^2+M_2^2}~~~~~\mbox{\rm and} ~~~~~
u_0 = \frac{M_1^2}{M_1^2+M_2^2}~,\nonumber
\end{eqnarray}
and $M_1^2$ and $M_2^2$ are the Borel parameters.
Using this expression and performing the integrations over the variables
$\alpha$ and $u$, we finally get
\begin{eqnarray}
\int du \varphi_\pi(u) \int d^4 x \frac{e^{i(p+qu)x}}{x^2} i
\sigma_{\alpha\beta} x_\alpha q_\beta = 8 i \sigma_{\alpha\beta}
p_\alpha q_\beta \varphi_\pi(u_0)~. \nonumber
\end{eqnarray}
All other terms can be calculated similarly and for
the theoretical part it follows from Eq. (9) that
\begin{eqnarray}
\lefteqn{
\Pi^{theor} =} \nonumber \\
&&- \alpha f_\pi \Bigg\{ \varphi_\pi(u_0) \Bigg[
(1 \pm 1) \frac{m_s}{8 \pi^2} M^4 f_1(s_0/M^2)
- \frac{1}{3} ( -\lambda \pm \sigma ) M^2 f_0(s_0/M^2)
+\frac{m_0^2}{12}( -\lambda\pm\sigma ) \Bigg] \nonumber \\
&& + \mu_\pi \varphi_\sigma(u_0) \Bigg[ (1 \mp 1) \frac{1}{24 \pi^2}
M^4 f_1(s_0/M^2) + (1 \pm 1) \frac{8}{2304}
\frac{\left< g^2 G^2 \right>}{\pi^2}
+ (1 \mp 1) \frac{1}{18} m_s \langle \bar s s \rangle
\pm \frac{1}{9} m_s \sigma \Bigg] \nonumber \\
&&+ \left[ g_1(u_0) + G_2(u_0) \right] \Bigg[
-(1 \pm 1) \frac{1}{\pi^2} m_s M^2 f_0(s_0/M^2) +
\frac{4}{3} \left( -\lambda \pm \sigma \right) \Bigg] \Bigg\}~,
\end{eqnarray}
where the function
\begin{eqnarray}
f_n(s_0/M^2)=1-e^{-s_0/M^2}\sum_{k=0}^n \frac{(s_0/M^2)^k}{k!}~, \nonumber
\end{eqnarray}
is the factor used to
subtract the continuum, which is modeled by the dispersion integral in the
region $s_1,~s_2 \ge s_0$, $s_0$ being the continuum threshold (obviously
the continuum thresholds for the $\Lambda$ and $\Sigma$ channels are
different). Since masses of $\Lambda$
and $\Sigma$ are very close to each other, we can choose
$M_1^2$ and $M_2^2$ to be equal to each other, i.e., $M_1^2 = M_2^2 =2 M^2$,
from which it follows that $u_0=1/2$.
Performing double Borel transformation over the variables $p^2$ and
$(p+q)^2$ in the physical part (5) and then equating the the obtained result
to Eq. (10), we get the sum rules for $g_{\pi\Lambda\Sigma}$ and
$g_{\pi\Sigma\Sigma}$ coupling constants
\begin{eqnarray}
g_{Y\Sigma^+\pi^-} \lambda_Y \lambda_{\Sigma^+} =
e^{m^2/M^2} \Pi^{theor}~,
\end{eqnarray}
where $m\approx m_\Lambda \approx m_\Sigma$.
From Eq. (11) it follows that in determining the strong coupling constants
$g_{\pi\Lambda\Sigma}$ and $g_{\pi\Sigma\Sigma}$ the experimentally
undetermined residues $\lambda_\Lambda$ and $\lambda_\Sigma$ need to be
eliminated from sum rules. The residues $\lambda_\Lambda$ and
$\lambda_\Sigma$ are determined from corresponding mass sum rules for the
$\lambda$ and $\Sigma$ hyperons \cite{R18,R20} as follows
\begin{eqnarray}
\left| \lambda_\Lambda \right|^2 e^{-m_\Lambda^2/M^2} 32 \pi^4 &=& M^6
f_2(s_0^\Lambda/M^2) + \frac{2}{3} a m_s (1 - 3 \gamma) M^2
f_0(s_0^\Lambda/M^2)\nonumber \\
&&+ b M^2 f_0(s_0^\Lambda/M^2) + \frac{4}{9}
a^2 (3+4 \gamma)~,\\
\left| \lambda_\Sigma \right|^2 e^{-m_\Sigma^2/M^2} 32 \pi^4 &=& M^6
f_2(s_0^\Sigma/M^2) - 2 a m_s (1+\gamma)M^2 f_0(s_0^\Sigma/M^2) \nonumber \\
&&+b M^2 f_0(s_0^\Sigma/M^2) + \frac{4}{3} a^2~,
\end{eqnarray}
where
\begin{eqnarray}
a &=& - 2 \pi^2 \langle \bar q q \rangle ~, \nonumber \\
b &=& \frac{\alpha_s \langle G^2 \rangle }{\pi} \simeq 0.012~GeV^4~, \nonumber \\
\gamma &=& \frac{\langle \bar s s \rangle}{\langle \bar q q \rangle} - 1
\simeq - 0.2~, \nonumber
\end{eqnarray}
and the functions $f_0(x),~f_1(x)$ are presented just after Eq. (10).
The ratio of the Eqs. (11), (12) and (13) gives
\begin{eqnarray}
g_{\pi\Lambda\Sigma} &=& e^{m^2/M^2} \frac{\Pi^{theor}}
{\lambda_\Lambda \lambda_\Sigma}~, \\
g_{\pi\Sigma\Sigma} &=& e^{m^2/M^2} \frac{\Pi^{theor}}
{\lambda_\Sigma^2}~.
\end{eqnarray}
The main reason why we consider the above--mentioned ratio
rather than the individual sum rules themselves (i.e., first determine
$\lambda_\Lambda$ and $\lambda_\Sigma$ independently from Eqs. (12) and (13)
and substitute their obtained values in Eq. (11)) is that the sum
rules obtained from these ratios are more stable as is similar to the baryon
mass sum rules case. In addition to that the
uncertainties coming from various parameters such as quark condensate,
$m_0^2$, continuum threshold $s_0$ and Borel parameter, are reduced.
\section{Numerical analysis}
Now we are ready to perform the numerical analysis. The main nonperturbative
input parameters in the sum rules (11) are the pion wave
functions. In our calculations we have used the set of wave functions
proposed in \cite{R6}. The explicit expressions of the wave functions are
\begin{eqnarray}
\varphi_\pi(u,\mu) &=& 6 u \bar u \left[ 1 + a_2(\mu) C_2^{3/2} (2 u -1 ) +
a_4(\mu) C_4^{3/2} (2 u -1 )\right]~, \nonumber \\
\varphi_\sigma(u,\mu) &=& 6 u \bar u \left[ 1 + C_2 \frac{3}{2}
\left[5(u-\bar u)^2 -1 \right] + C_4 \frac{15}{8}
\left[21(u-\bar u)^4 - 14 (u-\bar u)^2 + 1\right]\right]~, \nonumber \\
g_1(u,\mu) &=& \frac{5}{2} \delta^2(\mu)\bar u^2 u^2 +
\frac{1}{2} \varepsilon(\mu) \delta^2(\mu) \Bigg[ u \bar u
(2 + 13 u \bar u) \nonumber \\
&& + 10 u^3 \ln u \left( 2 - 3u + \frac{6}{5} u^2 \right) +
10 \bar u^3 \ln \bar u \left( 2 - 3\bar u +
\frac{6}{5}\bar u^2 \right)\Bigg]~, \nonumber \\
G_2(u,\mu) &=& \frac{5}{3} \delta^2(\mu) \bar u^2 u^2~,
\end{eqnarray}
where $\bar u = 1-u$, $C_2^{3/2}$ and $C_4^{3/2}$ are the Gegenbauer
polynomials defined as
\begin{eqnarray}
C_2^{3/2} (2 u -1 ) &=& \frac{3}{2} \left[ 5(2 u-1)^2 +1 \right]~,\nonumber \\
C_4^{3/2} (2 u -1 ) &=& \frac{15}{8} \left[ 21 (2 u -1 )^4 -
14 (2 u -1 )^2 +1 \right] ~,
\end{eqnarray}
and $a_2(\mu=0.5~GeV)=2/3$, $a_4(\mu=0.5~GeV)=0.43$. The parameters
$\delta^2(\mu=1~GeV)=0.2~GeV^2$ \cite{R21} and $\epsilon(\mu=1~GeV)=0.5$
\cite{R6}. Furthermore $f_\pi=0.132~GeV$, $\mu_\pi(\mu=1~GeV)=1.65$,
$\langle \bar s s \rangle = 0.8 \langle \bar q q \rangle$ and
$\langle \bar q q \rangle\vert_{\mu=1~GeV}=-(0.243)^3~GeV^3,~s_0 = s_0^\Lambda \simeq
s_0^\Sigma = (3.0 \pm 0.2)~GeV^2$. Moreover all further calculations
are performed at $u=u_0=1/2$.
Having fixed the input parameters, one
must find the range of values of $M^2$ over which the sum rule
is reliable. The lowest possible value of $M^2$ is determined by
the requirement that the terms proportional to the highest inverse power of
the Borel parameters stay reasonable small. The upper bound of $M^2$ is
determined by demanding that the continuum contribution is not too large.
The interval of $M^2$ which satisfies both conditions is
$1 ~GeV^2 < M^2 < 2.5~GeV^2$. The analysis of the sum rules (14) and (15) shows that
the best stability is achieved in the region of $M^2$,
$1.4~GeV^2 < M^2 < 1.8~GeV^2$. This leads to the following result for the
coupling constants
\begin{eqnarray}
g_{\pi\Lambda\Sigma} &=& 5.3 \pm 1.8 \nonumber \\
g_{\pi\Sigma\Sigma} &=& 12.5 \pm 4.5~,
\end{eqnarray}
in which the errors coming from the quark condensate (which varies in the
region $-(0.24~GeV)^3$ and $-(0.26~GeV)^3$), $m_0^2$ parameter
(which varies in the region $0.6~GeV^2$ and $1.4~GeV^2$), variation of the
Borel parameter and change of the continuum threshold $s_0$ are all taken
into consideration. Our calculation shows that the main error comes from
uncertainties of the quark condensate. The central values of the coupling
constants are obtained at $m_0^2=0.8~GeV^2$,
$\langle \bar q q \rangle = - (0.243~GeV)^3 $ and $s_0=3.0~GeV^2$.
At this point we would like to stress that the above--mentioned strong
coupling constant have been analyzed using the individual sum rules
themselves by first determining $\lambda_\Lambda$ and $\lambda_\Sigma$
independently from Eqs. (12) and (13) and substituting their obtained values
in Eq. (11). The results predicted by this approach are close to the ones
presented in Eq. (18), however our calculations show that
the ratio sum rules prediction is more stable and reliable.
Here it should be noted that since the phase of the coupling constants can not be
predicted by the sum rules, we take into consideration
the magnitudes of these coupling constants to be able to compare them with
the predictions of other approaches.
Let us now compare our results on the coupling constants
$g_{\pi\Lambda\Sigma}$ and $g_{\pi\Sigma\Sigma}$ with SU(3) symmetry
prediction. As is known, SU(3) symmetry predicts
\begin{eqnarray}
G_{\pi\Lambda\Sigma} &=& \frac{2}{\sqrt{3}} (1-\alpha)\, G_{NN\pi} ~,\nonumber \\
G_{\pi\Sigma\Sigma} &=& 2 \alpha \,G_{NN\pi} ~,
\end{eqnarray}
where $\alpha = F/(F+D)$ (see for example \cite{R22}). Exact SU(3) symmetric
analysis of pion--baryon coupling gives $F/D \simeq 0.58$
\cite{R22} (exact SU(6) symmetry predicts
$F/D=2/3$). It follows from Eqs. (19) that
\begin{eqnarray}
R = \frac{G_{\pi\Sigma\Sigma}}{G_{\pi\Lambda\Sigma}} =
\frac{\sqrt{3}\alpha}{1-\alpha} \approx 1~,
\end{eqnarray}
while our analysis yields $R\simeq 2$. It follows from a comparison of these
results that SU(3) symmetry is broken significantly.
The pion--baryon couplings predicted in \cite{R12} are not
presented here since we have already noted that the results of
this work is currently under debate in literature.
In conclusion we have calculated the strong coupling constants of pion with
$\Lambda$ and $\Sigma$ hyperons and found out that our results differ
significantly from that of the SU(3) symmetry prediction.
\newpage
|
\section{Introduction}
Let $ \cal D $ be the Lie algebra of regular differential operators on
${ \Bbb C } \setminus \{0\}$, and ${ \widehat {\cal D} }= { \cal D } + { \Bbb C } C$ be the central
extension of ${ \cal D }$. In the representation theory of the Lie
algebra $ \widehat {\cal D} $ the most important are the irreducible quasi-finite
highest weight modules. These modules were classified by V. Kac
and A. Radul in \cite{KR1}. In \cite{FKRW} was shown that the
language of vertex algebras is very useful in $ \widehat {\cal D} $--module
theory. In particular, on the irreducible vacuum $ \widehat {\cal D} $--module
$L(0,c, \widehat {\cal D} )$ exists a natural structure of a simple vertex algebra
(cf. \cite{FKRW}, \cite{K}). This vertex algebra is usually
denoted by $W_{1+\infty,c}$. The results from \cite{KR1} give that
the representation theory of the vertex algebra $W_{1+\infty,c}$
is nontrivial only in the case $c \in { \Bbb Z }$. When $N \in { \Bbb N }$,
the irreducible modules for $W_{1+\infty,N}$ were classified in
\cite{FKRW}.
The structure of the vertex algebra $W_{1+\infty,-N}$ is much
complicated. The representation theory of $W_{1+\infty,-N}$ was
began by Kac and Radul in \cite{KR2}. They realized
$W_{1+\infty,-N}$ as a vertex subalgebra of the vertex algebra
$V_N$ constructed from Weyl algebra $W_N$ (we recall this result
in the Section \ref{reckr}). They also constructed a large class
of irreducible $W_{1+\infty,-N}$--modules. The next step in this
direction was made by Wang in \cite{W1}, \cite{W2}. He considered
the case $c=-1$, and proved that $W_{1+\infty,-1}$ is isomorphic
to tensor product $W_{3,-2} \otimes H_0$, where $W_{3,-2}$ is a
vertex algebra associated with $W_3$--algebra with the central
charge $-2$ and $H_0$ is a Heisenberg vertex algebra. He also
classified all irreducible modules for $W_{3,-2}$ and
$W_{1+\infty,-1}$. The representations obtained in \cite{W2}
weren't identified as a highest weight $ \widehat {\cal D} $--modules.
In the present paper we will construct $2 N$--dimensional family
of irreducible $W_{1+\infty,-N}$--modules. Our family includes the
modules constructed in \cite{KR2}, and also coincides with the
modules from \cite{W2} in the case $c=-1$.
Let us explain our result in more details. We consider lattice
vertex superalgebra $V_L$ (cf. \ref{stwist}), and for suitably
chosen lattice $L$ we show that $V_N$ and $W_{1+\infty,-N}$ are
vertex subalgebras of $V_L$ (cf. Section \ref{wsec}). This fact is
in physical literature known as a bosonization of $\beta \gamma$
system (see \cite{FMS}, \cite{W1}). The embedding of
$W_{1+\infty,-N}$ into vertex superalgebra $V_L$ will imply that
$W_{1+\infty,-N}$ can be realized as a vertex subalgebra of the
Heisenberg vertex algebra $M(1)$ with $2 N$--generators. We
explicitly identify the generators of $W_{1+\infty,-N}$ in terms
of Schur polynomials. Considering $M(1)$--modules $M(1,\lambda)$ we
obtain $W_{1+\infty,-N}$--modules $V(\lambda,-N)$ as a irreducible
subquotients of $M(1,\lambda)$ (see Section \ref{wsec}). Considering
$V(\lambda,-N)$ as a $ \widehat {\cal D} $--module, we identify its highest weights.
As a result we get that all modules $V(\lambda,-N)$ are quasi-finite
$ \widehat {\cal D} $--modules.
\section{Vertex superalgebras}
In this section we recall the definition of vertex (super)algebra
and a few basic formulas. Fore more details about vertex
(super)algebras its representations, and representation theory of
certain examples of vertex (super)algebras see \cite{B},
\cite{DL}, \cite{FLM}, \cite{K}, \cite{Li}, \cite{A}.
Let $V$ be a vector space. A
{\it field} is a series of the form
$$
a (z) = \sum_{n \in { \Bbb Z } } a_n z^{- n - 1},
$$ where $a_n \in \mbox{End} V$ are such that for each $v \in V$
one has: $a_n v = 0 \quad \mbox{for} \quad n \gg 0$. Here $z$ is
a formal indeterminate.
For a ${ \Bbb Z } _2$-graded vector space $W=W^{even}+W^{odd}$ we write
$|u|\in { \Bbb Z }_2$, a degree of $u$, only for homogeneous elements:
$|u|=0$ for an even element $u\in W^0$ and $|u|=1$ for an odd
element $u\in W^1$. For any two $ \Bbb Z _2$-homogeneous elements $u$
and $v$ we define $\epsilon _{u,v}=(-1)^{|u||v|}\in { \Bbb Z }$.
\begin{definition}
A {\it vertex superalgebra} is a quadruple
$(V,{ \bf 1 },D,Y)$, where $V=V^{even}\oplus V^{odd}$ is a
$\Bbb{Z}_{2}$-graded vector space, $D$ is a
$\Bbb{Z}_{2}$-endomorphism of $V$, ${ \bf 1 }$ is a specified vector
called the {\it vacuum} of $V$, and $Y$ is a linear map
\begin{eqnarray} Y(\cdot,z):& &V\rightarrow ({\rm
End}V)[[z,z^{-1}]];\nonumber\\ & &a\mapsto Y(a,z)=\sum_{n\in
\Bbb{Z}}a_{n}z^{-n-1}\;\;(\mbox{where } a_{n}\in {\rm End}V)
\nonumber
\end{eqnarray}
such that
%
\begin{eqnarray} (V1)& &\mbox{For any }a,b\in V, a_{n}b=0\;\;\;\mbox{ for }n
\mbox{ sufficiently large;}\nonumber \\ (V2)&
&[D,Y(a,z)]=Y(D(a),z)=\partial_z Y(a,z)\;\;\mbox{ for any }a\in
V;\nonumber
\\ (V3)& &Y({ \bf 1 },z)=Id_{V}\;\;\;\mbox{(the identity operator of
$V$)};\nonumber \\ (V4)& &Y(a,z){ \bf 1 }\in ({\rm End}V)[[z]] \mbox{
and }\lim_{z \rightarrow 0}Y(a,z){ \bf 1 }=a\;\;\mbox{ for any }a\in
V;\nonumber \\ (V5)& &\mbox{For }\Bbb{Z}_{2}\mbox{ -homogeneous
}a,b\in V, \mbox{ the following {\it Jacobi identity} holds:}
\nonumber \end{eqnarray}
\begin{eqnarray} &
&\;\;\;z_{0}^{-1}\delta\left(\frac{z_{1}-z_{2}}{z_{0}}\right)Y(a,z_{1})Y(b,z_{2})
-\varepsilon_{a,b}z_{0}^{-1}\delta\left(\frac{z_{2}-z_{1}}{-z_{0}}\right)
Y(b,z_{2})Y(a,z_{1})\nonumber \\ &
&=z_{2}^{-1}\delta\left(\frac{z_{1}-z_{0}}{z_{2}}\right)Y(Y(a,z_{0})b,z_{2}).\nonumber
\end{eqnarray}
\end{definition}
In the case $V= V^{even}$, i.e. when all vectors are even we say
that $V$ is vertex algebra.
From the Jacobi identity follows
%
\begin{eqnarray} \label{comut}
\left[
a_{m}, b_{n}
\right] = \sum^\infty_{j = 0}
{ m \choose
j} (a_{j} b)_{m + n - j}, \quad m, n \in { \Bbb Z },
\end{eqnarray}
\begin{eqnarray} \label{5.2}
Y (a_{- 1} b, z) =
: Y(a, z) Y(b, z):,
\end{eqnarray}
%
where the normally ordered product is defined, as usual, by
%
$$
: Y (a, z) Y (b, z): = Y (a, z)_- Y(b, z) + Y (b, z)
Y (a, z)_+
$$
%
and
%
$$
Y(a, z)_+ = \sum_{j \in { \Bbb Z }_+} a_{j} z^{-j - 1},
\quad Y(a, z)_- = Y(a, z) - Y (a, z)_+
$$
%
are the {\it annihilation\/} and the {\it creation\/} parts of $Y
(a, z)$.
\section{ The vertex algebra $W_{1 + \infty, - N}$ }
\label{reckr}
In this section we recall some of the results from \cite{KR1},
\cite{KR2}, \cite{FKRW}.
Let $ \cal D $ be the Lie algebra of complex regular differential
operators on ${ \Bbb C }^\times$ with the usual bracket, in an
indeterminate $t$. The elements
%
$$
J^l (k) = - t^{l + k} (\partial_t)^l \quad
(k \in { \Bbb Z }, l \in { \Bbb Z }_+)
$$
%
form a basis of ${ \cal D }$. The Lie algebra ${ \cal D }$ has the following
2-cocycle with values in ${ \Bbb C }$ :
$$
\Psi (f (t) \partial_t^m, g (t) \partial_t^n) =
\frac{m! n!}{(m + n + 1)!}
\mbox{Res}_{t = 0} f^{(n + 1)} (t) g^{(m)} (t) dt,
$$
%
where $f^{(m)} (t) = \partial^m_t f (t)$. We denote by ${\widehat
{ \cal D } } = { \cal D } \oplus { \Bbb C } C$, where $C$ is the central element,
the corresponding central extension of the Lie algebra ${ \cal D }$.
Another important basis of ${ \cal D }$ is
%
$$
L^l (k) = - t^k D^l \quad
(k \in { \Bbb Z }, l \in { \Bbb Z }_+)
$$
%
where $D = t \partial_t$. These two bases are related by the
formula \cite{KR1}:
%
\begin{eqnarray} \label{veza}
J^l (k) = - t^k D (D - 1) \cdots (D - l + 1).
\end{eqnarray}
%
Given a sequence of complex numbers $\lambda = (\lambda_j)_{j
\in { \Bbb Z }_+}$ and a complex number $c$ there exists a unique
irreducible ${\widehat { \cal D } }$-module $L (\lambda, c ; {\widehat
\cal D })$ which admits a non-zero vector $v_\lambda\/$ such that:
%
$$
L^j (k) v_\lambda = 0 \quad \mbox{for} \quad
k > 0,\; L^j (0) v_\lambda = \lambda_j v_\lambda, \; C = c I.
$$
%
This is called a highest weight module over $\widehat { \cal D }$ with
highest weight $\lambda\/$ and central charge $c$. The module $L
(\lambda, c; \widehat \cal D )$ is called {\it quasifinite\/} if all
eigenspaces of ${D}$ are finite-dimensional (note that $D\/$ is
diagonalizable). It was proved in (\cite{KR1} Theorem 4.2) that
$L (\lambda, c; \hat D)$ is a quasi-finite module if and only if
the generating series
$$ \Delta_\lambda (x) = \sum^\infty_{n = 0} \frac{x^n}{n!}
\lambda_n $$
has the form $$
\Delta_\lambda (x) = \frac{\phi (x)}{e^x - 1},
$$ where
$$
\phi (x) + c = \sum_i p_i (x) e^{r_i x} \quad
\mbox{(a finite sum)},
$$
%
$p_i (x)$ are non-zero polynomials in $x\/$ such that $\sum_i p_i
(0) = c$ and $r_i\/$ are distinct complex numbers. The numbers
$r_i\/$ are called the {\it exponents\/} of this module and the
polynomials $p_i (x)$ are called their {\it multiplicities}.
Recall now that the ${ \widehat {\cal D} } $-module $L(0, c; {\hat { \cal D } })$ has a
canonical structure of a vertex algebra with the vacuum vector
${ \bf 1 } = v_0$ and generated by the fields $J^k (z) = \sum_{m \in
{ \Bbb Z } } J^k (m) z^{-m - k - 1}$ \cite{FKRW}.
In \cite{KR2}, the vertex algebra $L(0,-N, \cal D )$ was realized by
using Weyl algebra ${ W_N}$ and the corresponding vertex algebra
$V_N$.
The Weyl algebra ${ W_N}$ is an associative algebra over ${ \Bbb C }$
generated by $a_i (m), a^{*}_i (m)$ $ (i=1,\dots, N; m \in { \Bbb Z })$
and $C$ with the following defining relations
$$ [a_i (m), a_j (n)]= [a^{*}_i (m) , a^{*}_j (n) ] = 0, \quad
[a^{*}_i (m), a_j (n) ] = \delta_{i,j} \delta_{m,-n} C
$$
for all $i,j \in \{ 1,\dots, N\}$, $n,m \in { \Bbb Z }$,
and $C$ is central element.
The vacuum ${ W_N}$--module $V_N$ is generated by one vector
${ \bf 1 }$, and the following relations
$$ a_i (n) { \bf 1 } = 0, \quad n\ge 0; \quad
a^{*}_i (n) { \bf 1 }= 0, \quad n> 0, \quad C { \bf 1 }= { \bf 1 }.$$
Define the following fields acting on $V_N$.
%
\begin{eqnarray} \label{polja}
a_i (z ) = \sum_{n \in { \Bbb Z } } a(n) z^{-n-1}, \quad
a^{*} _i (z) = \sum_{n \in { \Bbb Z } } a^{*} (n) z^{-n}.
\end{eqnarray}
Then there is a unique extensions of the fields (\ref{polja})
such that $V_N$ becomes a vertex algebra (see \cite{KR2}).
Denote by $W_{1 + \infty, -N}$ a vertex subalgebra of $V_N$
generated by the fields
$$
J^k (z) = - \sum^N _{i = 1} : a_i (z) \partial{^k} _{z} a^{*}
_i (z): .
$$
%
\begin{proposition}(\cite{KR2}, Proposition 6.3)
We have an isomorphism of vertex algebras:
%
$$ L (0, -N; {\hat { \cal D }}) \simeq W_{1 + \infty, -N} $$
%
under which the fields (denoted by the same symbol) $J^k (z)$
correspond to each other.
\end{proposition}
\section{Lattice and Heisenberg vertex algebras } \label{stwist}
Let $L$ be a lattice.
Set ${ \frak h}={ \Bbb C }\otimes_{ \Bbb Z }L$ and
extend the ${ \Bbb Z }$-form $ \langle \cdot, \cdot \rangle $ on $L$ to ${ \frak h}$.
Let $\hat{{ \frak h}}={ \Bbb C }[t,t^{-1}]\otimes { \frak h} \oplus { \Bbb C }c$ be the affinization of
${ \frak h}.$ We also use the notation $h(n)=t^{n}\otimes h$ for $h\in
{ \frak h}, n\in { \Bbb Z }$.
Set
$
\hat{{ \frak h}}^{+}=t{ \Bbb C }[t]\otimes
{ \frak h};\;\;\hat{{ \frak h}}^{-}=t^{-1}{ \Bbb C }[t^{-1}]\otimes { \frak h}.
$
Then $\hat{{ \frak h}}^{+}$ and $\hat{{ \frak h}}^{-}$ are abelian subalgebras
of $\hat{{ \frak h}}$. Let $U(\hat{{ \frak h}}^{-})=S(\hat{{ \frak h}}^{-})$ be the
universal enveloping algebra of $\hat{{ \frak h}}^{-}$. Let ${\lambda} \in
{ \frak h}$. Consider the induced $\hat{{ \frak h}}$-module
\begin{eqnarray*}
M(1,{\lambda})=U(\hat{{ \frak h}})\otimes _{U({ \Bbb C }[t]\otimes { \frak h}\oplus
{ \Bbb C }c)}{ \Bbb C }_{\lambda}\simeq
S(\hat{{ \frak h}}^{-})\;\;\mbox{(linearly)},\end{eqnarray*} where
$t{ \Bbb C }[t]\otimes { \frak h}$ acts trivially on ${ \Bbb C }$,
${ \frak h}$ acting as $\langle h, {\lambda} \rangle$ for $h \in { \frak h}$
and $c$ acts on ${ \Bbb C }$ as multiplication by 1. We shall write
$M(1)$ for $M(1,0)$.
For $h\in { \frak h}$ and $n \in { \Bbb Z }$ write $h(n) = t^{n} \otimes h$. Set
$
h(z)=\sum _{n\in { \Bbb Z }}h(n)z^{-n-1}.
$
Then $M(1)$ is vertex algebra which is generated by the fields
$h(z)$, $h \in { \frak h}$, and $M(1,{\lambda})$, for $\lambda \in { \frak h}$, are
irreducible modules for $M(1)$.
Let $\hat{L}$ be the canonical central extension of $L$ by the
cyclic group $\< \pm 1\>$:
\begin{eqnarray}\label{2.7}
1\;\rightarrow \< \pm 1\>\;\rightarrow \hat{L}\;\bar{\rightarrow}
L\;\rightarrow 1
\end{eqnarray}
with the commutator map $c(\alpha,\beta)=(-1)^{\< \alpha,\beta\>}$
for $\alpha,\beta \in L$. Let $e: L\to \hat L$ be a section such
that $e_0=1$ and $\epsilon: L\times L\to \<\pm 1\>$ be the
corresponding 2-cocycle. Then
$\epsilon(\alpha ,\beta )\epsilon(\beta ,\alpha )=(-1)^{\<\alpha ,\beta \>},$
\begin{equation}\label{2c}
\epsilon (\alpha ,\beta )\epsilon (\alpha +\beta ,\gamma)=\epsilon (\beta ,\gamma)\epsilon (\alpha ,\beta +\gamma)
\end{equation}
and $e_{\alpha }e_{\beta }= \epsilon (\alpha ,\beta )e_{\alpha +\beta }$ for $\alpha ,\beta ,\gamma\in L.$
Form the induced $\hat{L}$-module
\begin{eqnarray*}
{ \Bbb C }\{L\}={ \Bbb C }[\hat{L}]\otimes _{\< \pm 1\>}{ \Bbb C }\simeq
{ \Bbb C }[L]\;\;\mbox{(linearly)},\end{eqnarray*} where ${ \Bbb C }[\cdot]$
denotes the group algebra and $-1$ acts on ${ \Bbb C }$ as
multiplication by $-1$. For $a\in \hat{L}$, write $\iota (a)$ for
$a\otimes 1$ in ${ \Bbb C }\{L\}$. Then the action of $\hat{L}$ on ${ \Bbb C }
\{L\}$ is given by: $a\cdot \iota (b)=\iota (ab)$ and $(-1)\cdot
\iota (b)=-\iota (b)$ for $a,b\in \hat{L}$.
Furthermore we define an action of ${ \frak h}$ on ${ \Bbb C }\{L\}$ by:
$h\cdot \iota (a)=\< h,\bar{a}\> \iota (a)$ for $h\in { \frak h},a\in
\hat{L}$. Define $z^{h}\cdot \iota (a)=z^{\< h,\bar{a}\> }\iota
(a)$.
The untwisted space associated with $L$ is defined to be
\begin{eqnarray*}
V_{L}={ \Bbb C }\{L\}\otimes _{{ \Bbb C }}M(1)\simeq { \Bbb C }[L]\otimes S(\hat{ \frak h}
^{-})\;\;\mbox{(linearly)}.
\end{eqnarray*} Then
$\hat{L},\hat{{ \frak h}},z^{h}\;(h\in { \frak h})$ act naturally on $V_{L}$
by acting on either ${ \Bbb C }\{L\}$ or $M(1)$ as indicated above.
Define ${ \bf 1 }= \iota ( e_0) \in V_L$.
We use a normal ordering procedure, indicated by open colons,
which signify that in the enclosed expression, all creation
operators $h(n)$ $(n<0)$,$a\in \hat{L}$ are to be placed to the
left of all annihilation operators $h(n),z^{h}\;(h\in { \frak h},n\ge
0)$. For $a \in \hat{L}$, set
\begin{eqnarray*}
Y(\iota (a),z)= : e^{\int
(\bar{a}(z)-\bar{a}(0)z^{-1})}az^{\bar{a}}:.
\end{eqnarray*}
Let $a\in \hat{L};\;h_{1},\cdots,h_{k}\in
{ \frak h};n_{1},\cdots,n_{k}\in { \Bbb Z }\;(n_{i}> 0)$. Set
\begin{eqnarray*}
v= \iota (a)\otimes h_{1}(-n_{1})\cdots h_{k}(-n_{k})\in
V_{L}.\end{eqnarray*}
Define vertex operator $Y(v,z)$ with
\begin{eqnarray} \label{defvertex} :\left({1\over (n_{1}-1)!}({d\over
dz})^{n_{1}-1}h_{1}(z)\right)\cdots \left({1\over
(n_{k}-1)!}({d\over dz})^{n_{k}-1}h_{k}(z)\right)Y(\iota (a),z): .
\end{eqnarray}
This gives us a well-defined linear map
\begin{eqnarray*}
Y(\cdot,z):& &V_{L}\rightarrow
(\mbox{End}V_{L})[[z,z^{-1}]]\nonumber\\ & &v\mapsto Y(v,z)=\sum
_{n\in { \Bbb Z }}v_{n}z^{-n-1},\;(v_{n}\in {\rm End}V_{L}).
\end{eqnarray*}
Let $\{\; h_{i}\;|\;i=1,\cdots,d\}$ be an orthonormal basis of
${ \frak h}$ and set
\begin{eqnarray*}
\omega ={1\over 2}\sum _{i=1}^{d} h_{i}(-1) h_{i}(-1)\in V_{L}.
\end{eqnarray*}
Then $Y(\omega,z)=\sum_{n\in { \Bbb Z }}L_n z^{-n-2}$ gives rise to a
representation of the Virasoro algebra on $V_{L}$ with the central
charged $d$ and
\begin{eqnarray} \label{vir.rel}
& &L_0\left(\iota(a)\otimes h_{1}(-n_{1})\cdots
h_{n}(-n_{k})\right)\nonumber \\ &=&\left({1\over 2}\<
\bar{a},\bar{a}\>+n_{1}+\cdots+n_{k}\right) \left(\iota(a)\otimes
h_{1}(-n_{1})\cdots h_{k}(-n_{k})\right).
\end{eqnarray}
The following theorem was proved in \cite{DL} and \cite{K}.
\begin{theorem}
$(V_L, { \bf 1 }, L_{-1}, Y)$ is vertex superalgebra.
\end{theorem}
Vertex algebra $M(1)$ can be treated as a subalgebra of $V_L$.
Define the Schur polynomials $p_{r}(x_{1},x_{2},\cdots)$ $(r\in
{ \Bbb Z }_{+})$ in variables $x_{1},x_{2},\cdots$ by the following
equation:
\begin{eqnarray}\label{eschurd}
\exp \left(\sum_{n= 1}^{\infty}\frac{x_{n}}{n}y^{n}\right)
=\sum_{r=0}^{\infty}p_{r}(x_1,x_2,\cdots)y^{r}.
\end{eqnarray}
For any monomial $x_{1}^{n_{1}}x_{2}^{n_{2}}\cdots x_{r}^{n_{r}}$
we have an element $h(-1)^{n_{1}}h(-2)^{n_{2}}\cdots
h(-r)^{n_{r}}{ \bf 1 }$ in both $M(1)$ and $V_{L}$ for $h\in{ \frak h}.$
Then for any polynomial $f(x_{1},x_{2}, \cdots)$, \\ $f(h(-1),
h(-2),\cdots){ \bf 1 }$ is a well-defined element in $M(1)$ and
$V_{L}$. In particular, $p_{r}(h(-1),h(-2),\cdots){ \bf 1 }$ for $r\in
{ \Bbb Z }_{+}$ are elements of $M(1)$ and $V_{L}$.
Suppose $a,b\in \hat{L}$ such that $\bar{a}=\alpha,\bar{b}=\beta$.
Then
\begin{eqnarray}\label{erelation}
Y(\iota(a),z)\iota(b)&=&z^{\<\alpha,\beta\>}\exp\left(\sum_{n=1}^{\infty}
\frac{\alpha(-n)}{n}z^{n}\right)\iota(ab)\nonumber\\
&=&\sum_{r=0}^{\infty}p_{r}(\alpha(-1),\alpha(-2),\cdots)\iota(ab)
z^{r+\<\alpha,\beta\>}.
\end{eqnarray}
Thus
\begin{eqnarray}\label{eab1}
\iota(a)_{i}\iota(b)=0\;\;\;\mbox{ for }i\ge -\<\alpha,\beta\>.
\end{eqnarray}
Especially, if $\<\alpha,\beta\>\ge 0$, we have
$\iota(a)_{i}\iota(b)=0$ for all $i\in { \Bbb Z }_{+}$, and if
$\<\alpha,\beta\>=-n<0$, we get
\begin{eqnarray}\label{eab}
\iota(a)_{i-1}\iota(b)=p_{n-i}(\alpha(-1),\alpha(-2),\cdots)\iota(ab)
\;\;\;\mbox{ for }i\in { \Bbb Z }_{+}.
\end{eqnarray}
We will need one structural result on Heisenberg vertex algebras.
Element $L_0$ of the Virasoro algebra defines a
${ \Bbb Z }_+$--graduation on vertex algebra $M(1)= \oplus_{n \in { \Bbb Z }_+
} M(1)_n$.
Let $v_{\lambda}$ be the highest weight vector in the
$M(1)$--module $M(1, \lambda)$.
The following lemma can be proved by using standard calculation in the
Heisenberg vertex algebras.
\begin{lemma} \label{standard}
Let $h \in { \frak h}$, and $n_1, \dots, n_r \in { \Bbb N }$. Let $k= n_1 +
\cdots + n_r$. Let $u= h (-n_1) \cdots h (-n_r) { \bf 1 }$, and
$ Y( u, z) = \sum_{n \in { \Bbb Z } } u_{n} z^{-n-1}$.
Then $u \in M(1)_k$, and we have
\begin{eqnarray} (1) && u_n v_{\lambda} = 0 \quad\mbox{for} \ n> k-1, \nonumber
\\
(2) && u_{k-1} v_{\lambda} = (-1) ^{n_1 + \cdots + n_r +r}( \langle \lambda, h
\rangle ) ^{r} v_{\lambda}. \nonumber \end{eqnarray}
\end{lemma}
\begin{proposition} \label{schur1}
Let $h \in { \frak h}$, and $r\in { \Bbb Z }_+$. Let $u =
p_{r}(h(-1),h(-2),\cdots){ \bf 1 } $. Set $ Y( u,z) = \sum _{n \in { \Bbb Z }
} u_n z^{-n-1}$. Then $u \in M_r$, and we have
\begin{eqnarray} (1) && u_n v_{\lambda} = 0 \quad \mbox{for} \ n>r-1, \nonumber \\
(2)&& u_{r-1} v_{\lambda} = { \langle \lambda, h \rangle \choose r} v_{\lambda}. \end{eqnarray}
\end{proposition}
{\em Proof.} Since $u \in M(1)_r$, we have that $u_n v_{\lambda} = 0$
for $n > r$. Set $x=\langle \lambda , h \rangle$. Using Lemma \ref{standard}
one can easily see that
$$ u_{r-1} v_{l}= p_r (x, -x, x,-x,\cdots ) v_{\lambda}.$$ Since
$$ \exp \left( \sum_{n=1} ^{\infty} \frac{(-1) ^{n-1} x}{n} y^{n}
\right) = ( 1+y) ^{x} = \sum_{r \ge 0} {x \choose r} y^{r},$$
we have that $ p_r (x, -x, x,-x,\cdots ) = {x \choose r}$, and we
conclude that $u_{r-1} v_{\lambda}= {x \choose r}$. \qed
\begin{remark} Proposition \ref{schur1} can be also proved by
using Zhu's algebra theory (see \cite{DLM}, Section 3.)
\end{remark}
\section{Representations of the vertex algebra $W_{1+\infty, -N}$ }
\label{wsec}
In this section we will construct $2N$--dimensional
family of irreducible modules for the vertex algebra $W_{1+\infty,
-N}$. We will prove that $W_{1+\infty, -N}$ can be realized as a
vertex subalgebra of Heisenberg vertex algebra $M(1)$.
First we will consider the following lattice:
\begin{eqnarray}
&& L=\bigoplus_{i= 1} ^{N} \left( { \Bbb Z } {\alpha }_i + { \Bbb Z } {\beta } _i
\right), \nonumber \quad \mbox{where}\\
&& \langle {\alpha }_i , {\alpha }_j \rangle = \delta_{i,j}, \quad \langle {\beta }_i ,
{\beta }_j \rangle = - \delta_{i,j}, \quad \langle {\alpha }_i, {\beta }_j \rangle = \langle
{\beta }_j , {\alpha }_i \rangle = 0, \nonumber \end{eqnarray}
for every $i,j \in \{1,\cdots,N\}$.
Let ${ \frak h}$, $M(1)$ and $V_L$
be defined as in the Section \ref{stwist}. Then ${ \frak h}$ is a
Heisenberg algebra, $\mbox{dim} { \frak h} = 2 N$, $M(1)$ is a
Heisenberg vertex algebra, and $V_L$ is a lattice vertex
superalgebra.
For every $ i\in \{1, \cdots,
N\}$, let $a^{i}, b^{i} \in {\hat L}$ such that
$${\overline{a^{i} } } = \alpha _i + \beta _i, \quad {\overline{b^{i} }}=-
(\alpha _i + \beta _i).$$
Then we define $e^{i},
f^{i} \in V_L$ with
$ e^{i} = \iota( a^{i} ), f^{i} = \iota( b^{i})$.
Set $Y( e^{i}, z) = \sum_{n \in { \Bbb Z } } e^{i} _n z^{-n-1}$, and
$Y( f^{i}, z) = \sum_{n \in { \Bbb Z } } f^{i} _n z^{-n-1}$.
Recall the definition of Schur polynomial $p_l ( h(-1),h(-2),
\dots)$ ($h\in { \frak h}$ ) from Section \ref{stwist}.
Set $p_l (h) := p_l ( h(-1),h(-2), \dots) \in S({\hat{ \frak h}} ^{}-)$.
Let $\gamma_i = \alpha _i + \beta _i$. Then $ \langle \gamma_i, \gamma_i \rangle
=0$.
Now, relations (\ref{eab1}) and (\ref{eab}) imply the following
lemma.
\begin{lemma} \label{fms1}
For all $i,j \in \{1, \cdots, N\}$, $k,n \in { \Bbb Z }$
we have :
\begin{eqnarray} (1) && e^{i} _n e^{j} = f^{i} _n f^{j}= 0, \quad \mbox{for}
\ \ n \ge 0, \nonumber \\
(2) && e^{i} _{-1} f^{j} = \delta_{i,j} { \bf 1 }, \quad f^{i} _{-1}
e^{j}= \delta_{i,j} { \bf 1 }, \nonumber \\
(3) && e^{i} _{-l-1} f^{i} = p_l( \gamma_i) { \bf 1 }, \quad f^{i} _{-l-1}
e^{i}= p_l (-\gamma_i) { \bf 1 }, \nonumber \\
(4)&& [ \alpha _i (k), e^{j} _n ]= \delta_{i,j} e^{j} _{n+k}, \quad [
\alpha _i (k), f^{j} _n ]= -\delta_{i,j} f^{j} _{n+k}. \end{eqnarray}
\end{lemma}
Define
\begin{eqnarray} &&A^{i} ( z) = \sum_{n \in { \Bbb Z } } A^{i} (n)
z^{-n-1}=Y(e^{i}\otimes \alpha_i (-1), z ) = \ :\alpha_i (z)
Y(e^{i} , z) :, \nonumber
\\
&& {{\bar A} }^{i} (z) = \sum_{n \in \Bbb Z } {{\bar A} } ^{i} (n) z^{-n}=
Y(f^{i} ,z) . \nonumber \end{eqnarray}
\begin{lemma} \label{fms2} We have
%
\begin{eqnarray}
&& [A^{i} (n), A ^{j} (m)]= [{{\bar A} } ^{i} (n), {{\bar A} } ^{j} (m)]= 0,
\nonumber \\
&& [{{\bar A} } ^{i} (n), A^{j} (m) ] = \delta_{i,j} \delta_{m+m,0},
\nonumber \end{eqnarray}
i.e. the components of the fields
$A^{i}(z), {{\bar A} } ^{i} (z)$, $1\le i \le N$, span an associative algebra isomorphic to the
Weyl algebra $W_N$ with $C=1$.
\end{lemma}
{\em Proof.}
Using (\ref{5.2}) we have that
$$ A^{i}(m) = \sum_{ n < 0} \alpha _i (n) e^{i }_{ m-n-1} + \sum_{n \ge
0} e^{i}_{m-n-1} \alpha _i (n) $$ for $1\le i \le N$.
Using Lemma \ref{fms1} we get
\begin{eqnarray}
&& A^{i} (0) f^{j}= -\delta_{i,j} f^{j}, \quad A^{i} (m) f^{j} = 0
\quad \mbox{for} \ m > 0, \nonumber \\
&& A^{i} (m) A^{j} (-1){ \bf 1 } = 0, \quad f^{i} _m f^{j} = 0, \quad
\mbox{for} \ m\ge 0. \nonumber \end{eqnarray}
Now, the statment of the lemma follows from the commutator
formulae (\ref{comut}). \qed
From Lemma \ref{fms2} follows the following result.
\begin{proposition} \label{wizom}The subalgebra of the vertex superalgebra $V_L$
generated by the fields $A^i (z)$, ${{\bar A} } ^{i} (z) $ $i=1, \dots,
N$, is isomorphic to the vertex algebra $V_N$. (Under this
isomorphism the fields $a_i (z) $ corresponds to $A^{i} (z)$, and
$a^{*} _i (z)$ to $\bar{A} ^{i}(z)$.)
\end{proposition}
{\em Proof.} From the Lemma \ref{fms2} follows that $V_L$ is a
module for the Weyl algebra ${ W_N}$. The subalgebra of $V_L$
generated by the fields $A^i (z)$ and ${{\bar A} } ^{i} (z) $ is exactly
the ${ W_N}$ submodule ${ W_N} . { \bf 1 }$ generated by ${ \bf 1 }$. Then we
have
$$ A^{i} (m) { \bf 1 } = {{\bar A} } ^{i} (n) { \bf 1 } = 0, \quad \mbox{for} \
m\ge0, \ n > 0.$$ Then the fact that $V_N$ is an irreducible
${ W_N}$--module implies
that $V_N \cong { W_N}. { \bf 1 } $. \qed
\begin{remark} In the physical literature the vertex algebra $V_N$
is known as $\beta \gamma$ system, and the lattice construction of
$\beta \gamma$ system is known as Friedan-Martinec-Shenker
bosonization (cf. \cite{FMS}, \cite{W1}). \end{remark}
With the respect to Proposition
\ref{wizom} we can identify vertex algebra $V_N$ with the
subalgebra of $V_L$ generated by the fields $A^{i}(z)$, $\bar{A}
^{i}(z)$. Since $W_{1 +\infty,-N}$ is a vertex subalgebra of $V_N$
generated by the fields
$$ J^{k} (z) = - \sum_{ i=1} ^{N} : A ^{i} (z)\partial_z ^{k}
{{\bar A} } ^{i} (z):$$
we have that $W_{1 +\infty,-N}$ is also a vertex subalgebra of
$V_L$. For every $i \in \{1, \cdots, N\}$, let $U^{i}_{k}= A^{i}
(-1)f^{i} _{-k-1} { \bf 1 } \in V_L$, and $U_{k}= \sum_{i=1} ^{N} U^{i}
_{k}$ . We have
$$ J^{k} (z) = -k! \sum_{i=1} ^{N } Y ( A^{i} (-1) {{\bar A} }^{i} (-k)
{ \bf 1 },z ) =-k! Y(U_k,z).$$
\begin{lemma} \label{ul1} We have
\begin{eqnarray}
&&U^{i} _{k} = A^{i}(-1) {{\bar A} }^{i} (-k) { \bf 1 }= \alpha _i (-1) p_k (-\gamma_i)
{ \bf 1 } + p_{k+1} (-\gamma_i) { \bf 1 }. \nonumber \end{eqnarray} In particular,
$U^{i}_{k} \in M(1)$.
\end{lemma}
{\em Proof.} Since $[A^{i} (-1), {{\bar A} }^{i} (-k)] = 0$, we have
that
$$
U^{i} _k = {{\bar A} }^{i} (-k)A^{i}(-1) { \bf 1 } = f^{i} _{-k-1} \alpha _i
(-1) e^{i}= \alpha _i (-1) f^{i}_{-k-1} e^{i} + f^{i} _{-k-2} e^{i}.$$
Then Lemma \ref{fms1} implies that
$$ U^{i} _{k} = \alpha _i (-1) p_k (-\gamma_i) { \bf 1 } + p_{k+1} (-\gamma_i) { \bf 1 },
$$ and lemma holds. \qed
\begin{theorem} Vertex algebra $W_{1+ \infty, -N}$ is a subalgebra of the vertex algebra $M(1)$.
\end{theorem}
{\em Proof.} From Lemma \ref{ul1} follows that $U_{k}= \sum_{i=1}
^{N} U^{i} _{k} \in M(1)$, for every $k \in { \Bbb Z }_+$. Since $J^{k}
(z) = - k! Y(U_{k},z)$, $k \in { \Bbb Z }_+$, generate $W_{1+ \infty,
-N}$, we have that $W_{1+ \infty, -N}$ is a subalgebra of the
vertex algebra $M(1)$. \qed
\begin{lemma} \label{hw}
For every $\lambda \in { \frak h}$, let $v_{\lambda}$ be the highest weight vector
in $M(1,\lambda)$. Then we have
$ J ^{k}(n) v_{\lambda} = 0$ for $ n > 0$, and
$$ J ^{k} (0) v_ {\lambda} = - k! \sum_{i=1} ^{N} \left( {-\langle \lambda , \gamma_i
\rangle \choose {k+1} } + \langle \lambda, \alpha _i \rangle {-\langle \lambda, \gamma_i \rangle \choose
k} \right) v_{\lambda} . $$
\end{lemma}
{\em Proof.} Set $Y( U^{i} _k, z) = \sum_{n \in { \Bbb Z } } U^{i} _k
(n) z^{-n-k-1}$. Proposition \ref{schur1}, and Lemma \ref{ul1}
implies that
\begin{eqnarray} \label{ul2}&& U^{i} _k (n) v_{\lambda} = 0 \quad \mbox{for} \ n
>0, \nonumber
\\ && U^{i} _k (0) v_{\lambda} = \langle \lambda, \alpha _i \rangle {-\langle \lambda , \gamma_i \rangle \choose k} +
{-\langle \lambda, \gamma_i \rangle \choose k+1}. \nonumber \end{eqnarray}
Since $ J^{k} (n) = - k!\sum_{i=1} ^{N} U^{i} _k (n)$, we have
that the
lemma holds. \qed
Since $W_{1+ \infty, -N}$ is a subalgebra of VOA $M(1)$,
then for every $\lambda \in { \frak h}$, $M(1, \lambda)$ is a $W_{1+ \infty,
-N}$--module. Let $V(\lambda,-N)$ be the irreducible $W_{1+ \infty,
-N}$--subquotient of $M(1, \lambda)$ generated by the vector $v_{\lambda}$.
\begin{theorem} \label{main}
For every $\lambda \in { \frak h} $ $ {V} ({\lambda},-N)$ is the irreducible
module for the vertex algebra $W_{1 +\infty, - N}$.
As a $ \widehat {\cal D} $--module, $V(\lambda,-N)$ is a irreducible quasi-finite
highest weight module, and the corresponding generating series is
\begin{eqnarray} \label{gener}
\Delta_{\lambda} (x) =-\sum_{i=1} ^{N}
\left(\frac{e^{ s_i x} -1 }{e^{x} -1} + t_i e^{s_i x} \right) ,
\end{eqnarray}
where
\begin{eqnarray} \label{defl}s_i = -\langle \lambda, \alpha _i + \beta _i \rangle, \quad t_i = \langle
\lambda, \alpha _i \rangle, \quad i=1,\dots, N. \end{eqnarray}
\end{theorem}
{\em Proof.} Lemma \ref{hw} implies that ${V}({\lambda},-N)$ is a
highest weight $ \widehat {\cal D} $--module. It remains to prove that ${
V}({\lambda},-N)$ is a quasi-finite $ \widehat {\cal D} $--module. In order to prove
this we have to identify the generating series $\Delta_{\lambda} (x)$.
Using the relation (\ref{veza}), it is straightforward to see the
following formulae.
\begin{eqnarray} \Delta_{\lambda} (x) = \sum_{k=0} ^{\infty} \frac{x^{k}}{k!}
L^{k} (0) v_{\lambda} = \sum_{k=0} ^{\infty} \frac{(e^{x} -1)^{k}}{k!}
J^{k} (0) v_{\lambda}. \nonumber \end{eqnarray}
Let $s_i, t_i$ be defined with (\ref{defl}). Then Lemma \ref{hw}
implies
\begin{eqnarray} \Delta_{\lambda} (x) = &&-\sum_{i=1} ^{N}\sum_{k=0} ^{\infty}
\left( {s_i \choose {k+1} } + t_i {s_i \choose k} \right) (e^{x}
-1)^{k} \nonumber
\\
= && -\sum_{i=1} ^{N} \left(\frac{e^{s_i x} -1 }{e^{x} -1} + t_i
e^{s_i x} \right) \nonumber
\\
=&& \frac{\Phi (x) }{e^{x} -1}, \nonumber \end{eqnarray}
where
$$ \Phi(x) -N = \sum_{i=1} ^{N} \left( (t_i -1 ) e^{s_i x} -t_i
e^{(s_i +1) x}\right).$$ This implies that ${V}(\lambda,-N)$ is a
quasi-finite $ \widehat {\cal D} $--module. \qed
\begin{remark}
Theorem \ref{main} gives the existence of $2 N$--dimensional
family of irreducible $W_{1+\infty,-N}$--modules. If we take in
(\ref{gener}) $t_i=0$ for every $i =1, \dots, N$, we get exactly
$W_{1+\infty,-N}$--modules constructed in \cite{KR2}.
\end{remark}
We
have the following conjecture.
\begin{conjecture} \label{slutnja}
The set $V(\lambda,-N)$, $\lambda \in { \frak h}$, lists all the irreducible
modules for the vertex algebra $W_{1+\infty,-N}$.
\end{conjecture}
\begin{remark} In Section \ref{sec-1} we will see that the Conjecture \ref{slutnja}
is true for $c=-1$. \end{remark}
\section{ The case of $c=-1$ } \label{sec-1}
In this section we will compare our results with the results from
\cite{W1}, \cite{W2}. In \cite{W1}, Wang proved that the vertex
algebra $W_{1+\infty,-1}$ is isomorphic to the tensor product
$W_{3,-2} \otimes H_0$, where $W_{3,-2}$ is a simple vertex
algebra associated to $W_3$--algebra with $c=-2$, and $H_0$ is a
Heisenberg vertex algebra. Moreover, in \cite{W2} Wang classified
all the irreducible modules for $W_{3,-2}$ and $W_{1+\infty,-1}$.
The methods used in \cite{W1}, \cite{W2} didn't imply the
identification of $W_{1+\infty,-1}$--modules as a highest weight
$ \widehat {\cal D} $--modules (see Section 5 in \cite{W2}). Our approach gives an
explicit identification of two-dimensional family
$W_{1+\infty,-1}$--modules in terms of highest weights.
Let $N=1$, and set $\alpha = \alpha _1$, $\beta = \beta _1$.
For $\lambda \in { \frak h}$ set $\lambda_{\alpha } = \langle \lambda, \alpha \rangle$, $\lambda_{\beta } = \langle
\lambda , \beta \rangle$. Let $M_{\alpha }(1, \lambda_{\beta })$ (resp. $M_{\beta } (1,
\lambda_{\beta })$ ) be the submodules of $M(1,\lambda)$ generated by the
highest weight vector $v_{\lambda}$ and $\alpha (n)$ (resp. $\beta (n)$). Set
$M_{\alpha } (1) = M_{\alpha }(1,0)$, $M_{\beta } (1) = M_{\beta }(1,0)$. Then
\begin{eqnarray} \label{m1ab} M(1) = M_{\alpha } (1)\otimes M_{\beta } (1), \quad
M(1,\lambda) = M_{\alpha } (1, \lambda_{\alpha } )\otimes M_{\beta } (1, \lambda _{\beta }). \end{eqnarray}
As in \cite{W2} we define
\begin{eqnarray}
&& T(z) = \frac{1}{2} \left( :\alpha (z) ^{2} : + \partial\alpha (z)
\right), \nonumber \\
&& W(z) = \frac{1}{12} \left( 4 : \alpha (z) ^{3} : + 6 : \alpha (z)
\partial\alpha (z) : + \partial^{2} \alpha (z) \right). \nonumber
\end{eqnarray}
\begin{theorem} \label{wang} \cite{W1}, \cite{W2}
\item[(1)] The fields $T(z)$, $W(z)$ span a subalgebra of $M_{\alpha }(1)$ isomorphic
to $W_{3,-2}$, and $W_{1+\infty,-1}\cong W_{3,-2} \otimes M_{\beta }
(1)$.
\item[(2)] Let ${\cal V}_r$ be the irreducible subquotient of
$W_{3,-2}$--module $M_{\alpha }(1,r)$. Then ${\cal V}_r$, $r \in { \Bbb C }$,
gives all the irreducible $W_{3,-2}$--modules.
\end{theorem}
Recall the definition of $W_{1+\infty,-1}$--modules $V(\lambda,-1)$
from Section \ref{wsec}. Then we have the following consequence of
Theorem \ref{main} and Theorem \ref{wang}.
\begin{corollary} We have
\item[(1)] $ {V}(\lambda,-1) \ = {\cal V}_{\lambda_{\alpha } } \otimes M_{\beta } (1,
\lambda_{\beta } )$ \ \ for every $\lambda \in { \frak h}$.
\item[(2)]The set $V(\lambda,-1)$, $\lambda \in { \frak h}$, gives all irreducible
$W_{1+\infty,-1}$--modules.
\item[(3)] $$\Delta_{\lambda} (x) =-\frac{e ^{-(\lambda_ {\alpha } + \lambda_{\beta }) x
}-1}{e^{x}-1}- \lambda_{\alpha } e^{- (\lambda_ {\alpha } + \lambda_{\beta }) x}.$$
\end{corollary}
{\em Proof.} (1) follows from the definition of ${\cal V}_{\alpha }$
and (\ref{m1ab}). Then theorem \ref{wang} implies that ${\cal
V}_r \otimes M_{\beta } (1,s)$, $r,s \in { \Bbb C }$, are all irreducible
$W_{1+\infty,-1}$--modules. Since
$ {V}(\lambda,-1) \ = {\cal V}_{\lambda_{\alpha } } \otimes M_{\beta } (1,
\lambda_{\beta } )$, we see that $V(\lambda,-1)$, $\lambda \in \frak h$ gives all
irreducible $W_{1+\infty,-1}$--modules, and we get (2). The
statement (3) follows from the Theorem \ref{main}. \qed
|
Subsets and Splits